Dopamine Handbook 0195373030, 9780195373035

The discovery of dopamine in 1957-1958 was one of the seminal events in the development of modern neuroscience, and has

411 104 15MB

English Pages 640 [632] Year 2009

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Dopamine Handbook
 0195373030, 9780195373035

Table of contents :
Preface
Contents
Contributors
1 Overview: A personal view of the dopamine neuron in historical perspective • Floyd E. Bloom
2 Neuroanatomy
2.1. Functional Neuroanatomy of Dopamine in the Striatum • Charles R. Gerfen
2.2. Functional Implications of Dopamine D2 Receptor Localization in Relation to Glutamate Neurons • Susan R. Sesack
2.3. Convergence of Limbic, Cognitive, and Motor Cortico-Striatal Circuits with Dopamine Pathways in Primate Brain • Suzanne N. Haber
2.4. The Relationship between Dopaminergic Axons and Glutamatergic Synapses in the Striatum: Structural Considerations • Jonathan Moss and J. Paul Bolam
3 Molecular pharmacology
3.1. Molecular Pharmacology of the Dopamine Receptors • Michele L. Rankin, Lisa A. Hazelwood, R. Benjamin Free, Yoon Namkung, Elizabeth B. Rex, Rebecca A. Roof, and David R. Sibley
3.2. Role of Dopamine Transporters in Neuronal Homeostasis • Marc G. Caron and Raul R. Gainetdinov
3.3. Intracellular Dopamine Signaling • Gilberto Fisone
3.4. Ion Channels and Regulation of Dopamine Neuron Activity • Birgit Liss and Jochen Roeper
4 Genes in development
4.1. Genetic Control of Meso-diencephalic Dopaminergic Neuron Development in Rodents • Wolfgang Wurst and Nilima Prakash
4.2. Factors Shaping Later Stages of Dopamine Neuron Development • Robert E. Burke
4.3. Postnatal Maturation of Dopamine Actions in the Prefrontal Cortex • Patricio O’Donnell and Kuei Y. Tseng
4.4. Genetic Dissection of Dopamine-Mediated Prefrontal-Striatal Mechanisms and Its Relationship to Schizophrenia • Hao-Yang Tan and Daniel R.Weinberger
5 Dopamine in prefrontal cortex and cognition
5.1. From Behavior to Cognition: Functions of Mesostriatal, Mesolimbic, and Mesocortical Dopamine Systems • Trevor W. Robbins
5.2. Contributions of Mesocorticolimbic Dopamine to Cognition and Executive Function • Stan B. Floresco
5.3. Dopamine’s Influence on Prefrontal Cortical Cognition: Actions and Circuits in Behaving Primates • Amy F.T. Arnsten, Susheel Vijayraghavan, Min Wang, Nao J. Gamo, and Constantinos D. Paspalas
5.4. Dopaminergic Modulation of Flexible Cognitive Control in Humans • Roshan Cools and Mark D’Esposito
5.5. Neurocomputational Analysis of Dopamine Function • Daniel Durstewitz
6 Striatum and midbrain—motor and motivational functions
6.1. Dopamine and Motor Function in Rat and Mouse Models of Parkinson’s Disease • Timothy Schallert and Sheila M. Fleming
6.2. Involvement of Nucleus Accumbens Dopamine in Behavioral Activation and Effort-Related Functions • John D. Salamone
6.3. Functional Heterogeneity in Striatal Subregions and Neurotransmitter Systems: Implications for Understanding the Neural Substrates Underlying Appetitive Motivation and Learning • Brian A. Baldo and Matthew E. Andrzejewski
6.4. Behavioral Functions of Dopamine Neurons • Philippe N. Tobler
7 Plasticity of forebrain dopamine systems
7.1. Dynamic Templates for Neuroplasticity in the Striatum • Ann M. Graybiel
7.2. Dopamine and Synaptic Plasticity in Mesolimbic Circuits • F. Woodward Hopf, Antonello Bonci, and Robert C. Malenka
7.3. Dopaminergic Modulation of Striatal Glutamatergic Signaling in Health and Parkinson’s Disease • D. James Surmeier, Michelle Day, Tracy S. Gertler, C . Savio Chan, and Weixing Shen
8 Dopamine mechanisms in addiction
8.1. The Role of Dopamine in the Motivational Vulnerability to Addiction • George F. Koob and Michel Le Moal
8.2. Dopaminergic Mechanisms in Drug-Seeking Habits and the Vulnerability to Drug Addiction • Barry J. Everitt, David Belin, Jeffrey W. Dalley, and Trevor W. Robbins
8.3. Imaging Dopamine’s Role in Drug Abuse and Addiction • Nora D. Volkow, Joanna S. Fowler, Gene-Jack Wang, Frank Telang, and Ruben Baler
9 Parkinson’s disease
9.1. Exploring the Myths about Parkinson’s Disease • Yves Agid and Andreas Hartmann
9.2. Pathophysiology of L-DOPA-Induced Dyskinesia in Parkinson’s Disease • M. Angela Cenci
9.3. Progression of Parkinson’s Disease Revealed by Imaging Studies • David J. Brooks
9.4. Transplantation of Dopamine Neurons: Extent and Mechanisms of Functional Recovery in Rodent Models of Parkinson’s Disease • Stephen B. Dunnett and Anders Björklund
9.5. Clinical Experiences with Dopamine Neuron Replacement in Parkinson’s Disease: What Is the Future? • Olle Lindvall
9.6. Novel Gene-Based Therapeutics Targeting the Dopaminergic System in Parkinson’s Disease • Deniz Kirik, Tomas Björklund, Shilpa Ramaswamy, and Jeffrey H. Kordower
9.7. Neuroprotective Strategies in Parkinson’s Disease • C. Warren Olanow
10 Schizophrenia and other psychiatric illnesses
10.1. Dopamine Dysfunction in Schizophrenia • Anissa Abi-Dargham, Mark Slifstein, Larry Kegeles, and Marc Laruelle
10.2. Neuropharmacological Profiles of Antipsychotic Drugs • Bryan L. Roth and Sarah C. Rogan
10.3. How Antipsychotics Work: Linking Receptors to Response • Nathalie Ginovart and Shitij Kapur
10.4. Dopamine Dysfunction in Schizophrenia: From Genetic Susceptibility to Cognitive Impairment • Heike Tost, Shabnam Hakimi and Andreas Meyer-Lindenberg
10.5. The Role of Dopamine in the Pathophysiology and Treatment of Major Depressive Disorder • Boadie W. Dunlop and Charles B. Nemeroff
10.6. Dopamine Modulation of Forebrain Pathways and the Pathophysiology of Psychiatric Disorders • Anthony A. Grace
Index

Citation preview

Dopamine Handbook

This page intentionally left blank

Dopamine Handbook Edited by

LESLIE L . IVERSEN, P H D, FRS Professor of Pharmacology University of Oxford Oxford, UK

SUSAN D. IVERSEN, P H D Department of Experimental Psychology University of Oxford Oxford, UK

STEPHEN B. DUNNET T, P H D School of Biosciences Cardiff University Cardiff, UK

ANDERS BJRKLUND, MD, P H D Wallenberg Neuroscience Center Division of Neurobiology Lund University Lund, Sweden

1

2010

1 Oxford University Press, Inc., publishes works that further Oxford University’s objective of excellence in research, scholarship, and education. Oxford New York Auckland Cape Town Dar es Salaam Hong Kong Karachi Kuala Lumpur Madrid Melbourne Mexico City Nairobi New Delhi Shanghai Taipei Toronto With offices in Argentina Austria Brazil Chile Czech Republic France Greece Guatemala Hungary Italy Japan Poland Portugal Singapore South Korea Switzerland Thailand Turkey Ukraine Vietnam

Copyright  2010 by Oxford University Press, Inc. Published by Oxford University Press, Inc. 198 Madison Avenue, New York, New York 10016 www.oup.com Oxford is a registered trademark of Oxford University Press. All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise, without the prior permission of Oxford University Press. Library of Congress Cataloging-in-Publication Data Dopamine handbook / [edited by] Leslie L. Iversen . . . [et al.]. p. ; cm. Includes bibliographical references and index. ISBN: 978-0-19-537303-5 1. Dopaminergic neurons—Handbooks, manuals, etc. 2. Dopaminergic mechanisms—Handbooks, manuals, etc. 3. Dopamine—Handbooks, manuals, etc. I. Iversen, Leslie L. [DNLM: 1. Dopamine—physiology—Handbooks. 2. Brain—physiology—Handbooks. 3. Dopamine Agents—Handbooks. 4. Receptors, Dopamine—Handbooks. WK 39 D692 2010] QP364.7.D666 2010 612.80 042—dc22 2009011010 9

8

7

6

5

4

3

2

1

Printed in the United States of America on acid-free paper

Preface

The discovery of dopamine as a neurotransmitter in the brain by Arvid Carlsson and colleagues in 1957 proved to be a seminal event in the development of modern neuroscience. Research on dopaminergic mechanisms in the past 50 years has been extremely productive in providing insights into such fundamental aspects of brain function as motor control, cognition, addiction, and reward. More than any other field of neurotransmitter research, dopamine research has provided important links between basic research and clinical practice––with the discovery of the first effective treatments for Parkinson’s disease, schizophrenia, and attention deficit hyperactivity disorder (ADHD) based on dopaminergic mechanisms. The Dopamine Handbook arose as an outcome of the symposium ‘‘Fifty Years of Dopamine Research’’ organized by Anders Bjo¨rklund and colleagues in Goteborg, Sweden, in 2007 to celebrate the 50th anniversary of Carlsson’s discovery. Although the proceedings gave rise to a series of excellent short articles in a special edition of Trends in Neuroscience (May 2007: Vol 30, No 5, pp. 185–250), we felt that the subject deserved a more detailed review. This culminated in the present handbook, with comprehensive reviews of all major topics in the dopaminergic field written by international experts. We are very grateful to the many contributors for undertaking the onerous task of writing lengthy reviews, and to the editors, Craig Panner and David D’Addona, at Oxford University Press for their patient overseeing of the project. Leslie Iversen Susan Iversen Stephen Dunnett Anders Bjo¨rklund

This page intentionally left blank

Contents

Contributors

xi

1

Overview: A personal view of the dopamine neuron in historical perspective Floyd E. Bloom

2

Neuroanatomy

3

4

3

9

2.1.

Functional Neuroanatomy of Dopamine in the Striatum Charles R. Gerfen

2.2.

Functional Implications of Dopamine D2 Receptor Localization in Relation to Glutamate Neurons 22 Susan R. Sesack

2.3.

Convergence of Limbic, Cognitive, and Motor Cortico-Striatal Circuits with Dopamine Pathways in Primate Brain 38 Suzanne N. Haber

2.4.

The Relationship between Dopaminergic Axons and Glutamatergic Synapses in the Striatum: Structural Considerations 49 Jonathan Moss and J. Paul Bolam

Molecular pharmacology

11

61

3.1.

Molecular Pharmacology of the Dopamine Receptors 63 Michele L. Rankin, Lisa A. Hazelwood, R. Benjamin Free, Yoon Namkung, Elizabeth B. Rex, Rebecca A. Roof, and David R. Sibley

3.2.

Role of Dopamine Transporters in Neuronal Homeostasis Marc G. Caron and Raul R. Gainetdinov

3.3.

Intracellular Dopamine Signaling Gilberto Fisone

3.4.

Ion Channels and Regulation of Dopamine Neuron Activity 118 Birgit Liss and Jochen Roeper

Genes in development

88

100

139

4.1.

Genetic Control of Meso-diencephalic Dopaminergic Neuron Development in Rodents 141 Wolfgang Wurst and Nilima Prakash

4.2.

Factors Shaping Later Stages of Dopamine Neuron Development Robert E. Burke

4.3.

Postnatal Maturation of Dopamine Actions in the Prefrontal Cortex Patricio O’Donnell and Kuei Y. Tseng

160 177

vii

viii

CONTENTS

4.4.

5

6

7

8

Genetic Dissection of Dopamine-Mediated Prefrontal-Striatal Mechanisms and Its Relationship to Schizophrenia 187 Hao-Yang Tan and Daniel R.Weinberger

Dopamine in prefrontal cortex and cognition

201

5.1.

From Behavior to Cognition: Functions of Mesostriatal, Mesolimbic, and Mesocortical Dopamine Systems 203 Trevor W. Robbins

5.2.

Contributions of Mesocorticolimbic Dopamine to Cognition and Executive Function Stan B. Floresco

5.3.

Dopamine’s Influence on Prefrontal Cortical Cognition: Actions and Circuits in Behaving Primates 230 Amy F.T. Arnsten, Susheel Vijayraghavan, Min Wang, Nao J. Gamo, and Constantinos D. Paspalas

5.4.

Dopaminergic Modulation of Flexible Cognitive Control in Humans Roshan Cools and Mark D’Esposito

5.5.

Neurocomputational Analysis of Dopamine Function Daniel Durstewitz

Striatum and midbrain—motor and motivational functions

215

249

261

277

6.1.

Dopamine and Motor Function in Rat and Mouse Models of Parkinson’s Disease Timothy Schallert and Sheila M. Fleming

6.2.

Involvement of Nucleus Accumbens Dopamine in Behavioral Activation and Effort-Related Functions 286 John D. Salamone

6.3.

Functional Heterogeneity in Striatal Subregions and Neurotransmitter Systems: Implications for Understanding the Neural Substrates Underlying Appetitive Motivation and Learning 301 Brian A. Baldo and Matthew E. Andrzejewski

6.4.

Behavioral Functions of Dopamine Neurons Philippe N. Tobler

Plasticity of forebrain dopamine systems

279

316

331

7.1.

Dynamic Templates for Neuroplasticity in the Striatum Ann M. Graybiel

7.2.

Dopamine and Synaptic Plasticity in Mesolimbic Circuits 339 F. Woodward Hopf, Antonello Bonci, and Robert C. Malenka

7.3.

Dopaminergic Modulation of Striatal Glutamatergic Signaling in Health and Parkinson’s Disease 349 D. James Surmeier, Michelle Day, Tracy S. Gertler, C . Savio Chan, and Weixing Shen

Dopamine mechanisms in addiction

333

369

8.1.

The Role of Dopamine in the Motivational Vulnerability to Addiction George F. Koob and Michel Le Moal

371

8.2.

Dopaminergic Mechanisms in Drug-Seeking Habits and the Vulnerability to Drug Addiction Barry J. Everitt, David Belin, Jeffrey W. Dalley, and Trevor W. Robbins

8.3.

Imaging Dopamine’s Role in Drug Abuse and Addiction 407 Nora D. Volkow, Joanna S. Fowler, Gene-Jack Wang, Frank Telang, and Ruben Baler

389

CONTENTS

9

10

Parkinson’s disease

419

9.1.

Exploring the Myths about Parkinson’s Disease Yves Agid and Andreas Hartmann

9.2.

Pathophysiology of L-DOPA-Induced Dyskinesia in Parkinson’s Disease M. Angela Cenci

9.3.

Progression of Parkinson’s Disease Revealed by Imaging Studies David J. Brooks

9.4.

Transplantation of Dopamine Neurons: Extent and Mechanisms of Functional Recovery in Rodent Models of Parkinson’s Disease 454 Stephen B. Dunnett and Anders Bjo¨rklund

9.5.

Clinical Experiences with Dopamine Neuron Replacement in Parkinson’s Disease: What Is the Future? 478 Olle Lindvall

9.6.

Novel Gene-Based Therapeutics Targeting the Dopaminergic System in Parkinson’s Disease Deniz Kirik, Tomas Bjo¨rklund, Shilpa Ramaswamy, and Jeffrey H. Kordower

9.7.

Neuroprotective Strategies in Parkinson’s Disease C. Warren Olanow

Schizophrenia and other psychiatric illnesses

421 434

445

498

509

10.1.

Dopamine Dysfunction in Schizophrenia 511 Anissa Abi-Dargham, Mark Slifstein, Larry Kegeles, and Marc Laruelle

10.2.

Neuropharmacological Profiles of Antipsychotic Drugs Bryan L. Roth and Sarah C. Rogan

10.3.

How Antipsychotics Work: Linking Receptors to Response Nathalie Ginovart and Shitij Kapur

10.4.

Dopamine Dysfunction in Schizophrenia: From Genetic Susceptibility to Cognitive Impairment 558 Heike Tost, Shabnam Hakimi and Andreas Meyer-Lindenberg

10.5.

The Role of Dopamine in the Pathophysiology and Treatment of Major Depressive Disorder 572 Boadie W. Dunlop and Charles B. Nemeroff

10.6.

Dopamine Modulation of Forebrain Pathways and the Pathophysiology of Psychiatric Disorders 590 Anthony A. Grace

Index

599

520 540

489

ix

This page intentionally left blank

Contributors

Anissa Abi-Dargham, MD Department of Psychiatry Columbia University and New York State Psychiatric Institute New York, NY Yves Agid, MD, PhD Centre d’Investigation Clinique Hoˆpital de la Pitie´-Salpeˆtrie`re Paris, France Matthew E. Andrzejewski, PhD Waisman Center University of Wisconsin-Madison Madison, WI Amy F.T. Arnsten, PhD Department of Neurobiology Yale Medical School New Haven, CT Brian A. Baldo, PhD Department of Psychiatry University of Wisconsin-Madison School of Medicine and Public Health Madison, WI Ruben Baler, PhD Office of Science Policy and Communications National Institute on Drug Abuse National Institutes of Health Bethesda, MD David Belin, PhD Poˆle Biologie Sante´ CNRS UMR 6187 & Universite´ de Poitiers 40, avenue du Recteur Pineau 86022 POITIERS cedex – FRANCE

Anders Bjo¨rklund, MD, PhD Wallenberg Neuroscience Center Division of Neurobiology Lund University Lund, Sweden Tomas Bjo¨rklund, M. Med Department of Experimental Medical Science Lund University Lund, Sweden Floyd E. Bloom, MD Molecular and Integrative Neuroscience Department The Scripps Research Institute La Jolla, CA J. Paul Bolam, PhD Medical Research Council, Anatomical Neuropharmacology Unit Department of Pharmacology University of Oxford Oxford, UK Antonello Bonci, MD Ernest Gallo Clinic and Research Center Department of Neurology Wheeler Center for the Neurobiology of Addiction Program in Neuroscience, University of California, San Francisco San Francisco, CA David J. Brooks, MD, DSc, FRCP, FMedSci Division of Neurosciences and Mental Health Imperial College London, UK Robert E. Burke, MD Department of Neurology The College of Physicians and Surgeons Columbia University New York, NY

xi

xii

CONTRIBUTORS

Marc G. Caron, PhD Department of Cell Biology Duke University Medical Center Durham, NC M. Angela Cenci, MD, PhD Department of Experimental Medical Science Lund University Lund, Sweden C. Savio Chan, PhD Department of Physiology Feinberg School of Medicine Northwestern University Chicago, IL

Daniel Durstewitz, PhD Computational Neuroscience Group Central Institute of Mental Health University of Heidelberg Mannheim, Germany Barry J. Everitt, ScD Department of Experimental Psychology Behavioral and Clinical Neuroscience Institute University of Cambridge Cambridge, UK Gilberto Fisone, PhD Department of Neuroscience Karolinska Institutet Stockholm, Sweden

Roshan Cools, PhD Donders Institute for Brain, Cognition and Behavior Radboud University Nijmegen Medical Centre Centre for Cognitive Neuroimaging Nijmegen, The Netherlands

Sheila M. Fleming, PhD Assistant Professor Departments of Psychology and Neurology University of Cincinnati Cincinnati, OH 45221

Jeffrey W. Dalley, PhD Department of Psychiatry Behavioral and Clinical Neuroscience Institute University of Cambridge Cambridge, UK

Stan B. Floresco, PhD Department of Psychology Brain Research Centre University of British Columbia Vancouver, B.C. Canada

Michelle Day, PhD Department of Physiology Feinberg School of Medicine Northwestern University Chicago, IL Mark D’Esposito, MD Helen Willis Neuroscience Institute University of California, Berkeley Berkeley, CA Boadie W. Dunlop, MD Department of Psychiatry and Behavioral Sciences Emory University School of Medicine Atlanta, GA Stephen B. Dunnett, PhD School of Biosciences Cardiff University Cardiff, UK

Joanna S. Fowler, PhD National Institute on Drug Abuse National Institutes of Health Bethesda, MD R. Benjamin Free, PhD Molecular Neuropharmacology Section National Institute of Neurological Disorders and Stroke National Institutes of Health Bethesda, MD Raul R. Gainetdinov, MD, PhD Department of Neuroscience and Brain Technologies Italian Institute of Technology Genova, Italy Nao J. Gamo Department of Neurobiology Yale Medical School New Haven, CT

CONTRIBUTORS

Charles R. Gerfen, PhD Laboratory of Systems Neuroscience National Institute of Mental Health National Institutes of Health Bethesda, MD Tracy S. Gertler, BS Department of Physiology Feinberg School of Medicine Northwestern University Chicago, IL Nathalie Ginovart, PhD Neuroimaging Unit Department of Psychiatry University of Geneva Geneva, Switzerland Anthony A. Grace, PhD Departments of Neuroscience, Psychiatry, and Psychology University of Pittsburgh Pittsburgh, PA Ann M. Graybiel, PhD Department of Brain and Cognitive Sciences McGovern Institute for Brain Research Massachusetts Institute of Technology Cambridge, MA Suzanne N. Haber, PhD Department of Pharmacology and Physiology University of Rochester School of Medicine Rochester, NY Shabnam Hakimi, BA Clinical Brain Disorders Branch National Institute of Mental Health National Institutes of Health Bethesda, MD Andreas Hartmann, MD, PhD Centre d’Investigation Clinique Hoˆpital de la Pitie´-Salpeˆtrie`re Paris, France Lisa A. Hazelwood, PhD Molecular Neuropharmacology Section National Institute of Neurological Disorders and Stroke National Institutes of Health Bethesda, MD

F. Woodward Hopf, PhD Ernest Gallo Clinic and Research Center Department of Neurology Wheeler Center for the Neurobiology of Addiction University of California, San Francisco San Francisco, CA Shitij Kapur, MD, PhD, FRCPC Institute of Psychiatry King’s College London London, UK Larry Kegeles, MD Departments of Psychiatry and Radiology Columbia University New York State Psychiatric Institute New York, NY Deniz Kirik, MD, PhD Department of Experimental Medical Science Lund University Lund, Sweden George F. Koob, PhD Committee of the Neurobiology of Addictive Disorders The Scripps Research Institute La Jolla, CA Jeffrey H. Kordower, PhD Department of Neurological Sciences Rush University Medical Center Chicago, IL Marc Laruelle, MD Schizophrenia and Cognitive Disorder Discovery Performance Unit Neurosciences Center of Excellence in Drug Discovery GlaxoSmithKline Harlow, UK Michel Le Moal, MD, PhD Neurocentre Magendie Institut National de la Sante´ et de la Recherche Me´dicale, Unite 862 Institut Franc¸ois Magendie Universite´ Victor Se´galen Bordeaux 2 Bordeaux, France

xiii

xiv

CONTRIBUTORS

Olle Lindvall, MD, PhD Laboratory of Neurogenesis and Cell Therapy Section of Restorative Neurology Wallenberg Neuroscience Center Lund University Hospital Lund, Sweden Birgit Liss, PhD Section Molecular Neurophysiology Department of General Physiology University of Ulm Ulm, Germany Robert C. Malenka, MD, PhD Nancy Pritzker Laboratory Department of Psychiatry and Behavioral Sciences co-Director, Stanford Institute for Neuro-Innovation and Translational Neurosciences Stanford University School of Medicine Palo Alto, CA Andreas Meyer-Lindenberg, MD, PhD Department of Psychiatry and Psychotherapy Central Institute of Mental Health University of Heidelberg Mannheim, Germany Jonathan Moss Medical Research Council, Anatomical Neuropharmacology Unit Department of Pharmacology University of Oxford Oxford, UK

C. Warren Olanow, MD, FRCPC Department of Neurology and Neuroscience Mount Sinai School of Medicine New York, NY Constantinos D. Paspalas, PhD Division of Neuroanatomy University of Crete School of Medicine Heraklion, Greece Nilima Prakash, PhD Institute of Developmental Genetics Helmholtz Center Munich German Research Center for Environmental Health Technical University Munich Munich/Neuherberg, Germany Shilpa Ramaswamy Department of Neurological Sciences Rush University Medical Center Chicago, IL Michele L. Rankin, PhD Molecular Neuropharmacology Section National Institute of Neurological Disorders and Stroke National Institutes of Health Bethesda, MD Elizabeth B. Rex, PhD Molecular Neuropharmacology Section National Institute of Neurological Disorders and Stroke National Institutes of Health Bethesda, MD

Yoon Namkung, PhD Molecular Neuropharmacology Section National Institute of Neurological Disorders and Stroke National Institutes of Health Bethesda, MD

Trevor W. Robbins, PhD Department of Experimental Psychology and Behavioral and Clinical Neuroscience Institute University of Cambridge Cambridge, UK

Charles B. Nemeroff, MD, PhD Department of Psychiatry and Behavioral Sciences Emory University School of Medicine Atlanta, GA

Jochen Roeper, MD, PhD Institute of Neurophysiology Neuroscience Center Goethe-University Frankfurt Frankfurt, Germany

Patricio O’Donnell, MD, PhD Department of Anatomy and Neurobiology University of Maryland School of Medicine Baltimore, MD

Sarah C. Rogan Department of Pharmacology University of North Carolina, Chapel Hill Chapel Hill, NC

CONTRIBUTORS

Rebecca A. Roof, PhD Molecular Neuropharmacology Section National Institute of Neurological Disorders and Stroke National Institutes of Health Bethesda, MD Bryan L. Roth, MD, PhD Departments of Pharmacology, Medicinal Chemistry, and Psychiatry Linberger Comprehensive Cancer Center and Neuroscience Center University of North Carolina, Chapel Hill Chapel Hill, NC John D. Salamone, PhD Division of Behavioral Neuroscience Department of Psychology University of Connecticut Storrs, CT Timothy Schallert, PhD Department of Psychology University of Texas at Austin Austin, TX Susan R. Sesack, PhD Departments of Neuroscience and Psychiatry University of Pittsburgh Pittsburgh, PA Weixing Shen, PhD Department of Physiology Feinberg School of Medicine Northwestern University Chicago, IL David R. Sibley, PhD Molecular Neuropharmacology Section National Institute of Neurological Disorders and Stroke National Institutes of Health Bethesda, MD Mark Slifstein, PhD Departments of Psychiatry and Radiology Columbia University New York State Psychiatric Institute New York, NY

D. James Surmeier, PhD Department of Physiology Feinberg School of Medicine Northwestern University Chicago, IL Hao-Yang Tan, MBBS, MMed, FRCPC Clinical Brain Disorders Branch National Institute of Mental Health National Institutes of Health Bethesda, MD Frank Telang, MD Brookhaven National Laboratory Upton, NY Philippe N. Tobler, PhD Department of Physiology, Development, and Neuroscience University of Cambridge Cambridge, UK Heike Tost, MD, PhD Clinical Brain Disorders Branch National Institute of Mental Health National Institutes of Health Bethesda, MD Kuei Y. Tseng, MD, PhD Department of Cellular and Molecular Pharmacology Rosalind Franklin University of Medicine and Science The Chicago Medical School Chicago, IL Susheel Vijayraghavan, PhD Laboratory of Neuropsychology National Institute of Mental Health National Institutes of Health Bethesda, MD Nora D. Volkow, MD Director, National Institute on Drug Abuse National Institutes of Health Bethesda, MD Gene-Jack Wang, MD Brookhaven National Laboratory Upton, NY

xv

xvi

CONTRIBUTORS

Min Wang, PhD Department of Neurobiology Yale Medical School New Haven, CT Daniel R. Weinberger, MD Clinical Brain Disorders Branch National Institute of Mental Health National Institutes of Health Bethesda, MD

Wolfgang Wurst, PhD Institute of Developmental Genetics Helmholtz Center Munich German Research Center for Environmental Health Technical University Munich Max Planck Institute of Psychiatry Munich, Germany

Dopamine Handbook

This page intentionally left blank

1 Overview: A personal view of the dopamine neuron in historical perspective F LOYD E . B LO OM

INTRODUCTION My goals for this overview perspective are to portray what I consider to have been the major discoveries in dopamine (DA) research in the central nervous system from the anatomical, synaptic, and neurohistochemical perspectives, in keeping with my own didactic hypothesis that ‘‘the gains in brain are mainly in the stain.’’1–5 This handbook was developed in the postcelebratory interval following the semicentennial of the discoveries that led to the recognition of DA as a neurotransmitter in its own right, independent of the other central catecholamines, norepinephrine and epinephrine. My historical perspective began with the first comprehensive review of the structure and function of the central DA neuronal systems done after the first 20 years of DA research.6 What stands out now, even more profoundly than it did in 1978, was how primitive our concepts were then of the range of regulatory functions that could be carried out by the systems of neurons characterized primarily on the basis of employing DA as their primary neurotransmitter (recall that 30 years ago, there was no thought that neurons might use more than one substance to transmit their interneuronal signals). Indeed, it was the ability to localize—neurocytologically—the neurotransmitters of the monoamine families that permitted what now can be seen as the progressive analysis of the synthesis, storage, release, conservation, and catabolism that has in consequence pioneered the functional conceptualizations of synaptic transmitter metabolism, not to mention the modes of action of

most psychotropic drugs. A large proportion of those early studies were done biochemically on samples of brain tissues dissected without regard to the characterizations of the neuronal circuits that made, stored, and released DA, and partly as a result, our understanding of the structure and function of the DA-containing cells and their circuits in the central nervous system emerged much more slowly.

HOW DO WE ‘ SEE’’ DA? Research into brain DA systems was accelerated because of the role attributed to this transmitter’s being deficient in Parkinson’s disease (see Agid and Hartmann, Chapter 9.1, this volume), acting excessively in schizophrenia (see Abi-Dargham et al., Chapter 10.1, this volume) and mediating the internal reward systems of the brain (see Koob and Le Moal, Chapter 8.1, this volume). Thus, it was recognized early in the course of this work that DA neuron systems are more complex in their anatomy, more diverse in localization and apparent function, and more numerous, both in terms of definable functional systems (i.e., motor, sensory and reward) and in numbers of neurons, than the other central catecholamine systems. The initial results from formaldehyde-induced fluorescence microscopy indicated that the DA neuron systems were principally located in the upper mesencephalon and diencephalon.7–10 These systems were not immediately accepted by the professional neuroanatomists of the time because the DA fiber 3

4

OVERVIEW

systems were fine, unmyelinated, and largely invisible to the degenerative tract tracing methods then in vogue and because the freeze-dry process that was required to drive the histochemical coupling with formaldehyde fragmented the brains. These problems were later overcome with the development of the vibratome-glyoxylic methods11,12 and by other molecular probes of the enzymes, receptors, and transporters (see below). In its early stages, the vast majority of DA research focus was on the major projection from the substantia nigra, the pars compacta to the neostriatum, and, somewhat later, to the projections between the ventral tegmental area and the olfactory tubercle and ventral striatum.6 With the advent of new, powerful neuroanatomical methods including optical13,14 and ultrastructural immunocytochemistry for synthetic enzymes,15 receptors, and transporters (see Sibley et al., Chapter 3.1, and Caron and Gainetdinov, Chapter 3.2, this volume), there has been a very rapid, marked increase in our understanding of the extent and organization of DA neuron systems. This understanding prominently included the detection of bone fide synaptic specializations when the ultrastructurally defined DA terminals in cortex and other forebrain target areas were analyzed with serial section reconstructions.16,17 Subsequent work has demonstrated DA neurons within the retina,18 within the olfactory bulb,19,20 and in projections to the hair cells in the cochlea.21,22 Not explicitly included in this handbook are the roles of hypothalamic dopaminergic systems in neuroendocrine regulation (see 23 for a recent review). The surprise finding of the early work was the demonstration of a small but concentrated projection of DA fibers to the rat prefrontal cortex. 24–26 As the compelling involvement of DA in human brain diseases became evident, repetition of the DA molecular mapping tools to the human and nonhuman primate cortex revealed a major, regionally and laminarly selective distribution of DA fibers in the primate medial prefrontal cortex27,28 that was far more prominent than that observed in the rodent’s limited frontal cortical area. Comparisons of the nigrostriatal synaptic arrangements of receptors and transporters with those of the cortical DA terminal fields revealed a striking difference: in striatum, based on the ultrastructural localization of the DA transporter, reuptake seems to occur close to the presumptive synaptic sites bearing the postsynaptic receptors, while in cortex, the transporters are quite remote from presumptive response sites, raising the reality of the popular theory of localized volume transmission (see Sibley et al., Chapter 3.1, and Caron and Gainetdinov, Chapter 3.2, this volume; but also see

below and 29). In addition, when DA neurons projecting to cortical target areas were characterized physiologically and neurochemically, it was observed that the cortically projecting neurons exhibited more burst firing than those projecting to the striatum, that their turnover of DA was faster and critically dependent on extracellular tyrosine, that these neuronal somata were less responsive to DA agonists and antagonists, and that in contrast to those projecting to striatum, they showed no tolerance to antipsychotic drugs.30–32 Those observations set the stages for the stunning functional depictions of the cellular effects of DA within the primate cortical circuitry that have been linked to cognition (see Robbins, Chapter 5.1; Floresco, Chapter 5.2; and Arnsten et al., Chapter 5.3, this volume) and provided hypotheses of the sequences of pathophysiology in schizophrenia33–35 (also see Meyer-Lindenberg et al., Chapter 10.4, this volume). Thus, were one forced to pick one thread that has empowered the dynamic thrust of DA research and its relevance for human disease, one need go no further than the ability to localize DA neurons and their circuits. While this handbook provides a remarkably complete analysis of the many active facets of DA research today, I want to concentrate my analysis of the question I first asked in my brief neurophysiological studies of DA, namely, what does DA do? It is important to note that the relatively slower rate of progress in elucidating the physiological properties of the DA neuron resulted in part from the fact that methods for the analysis of any catecholamine-containing pathway required techniques not previously available because at the time there were no other neurochemically defined neuronal pathways to be studied, and in part from the fact that once these techniques were available and in widespread use, the body of physiological data that was produced indicated that catecholamine neuron systems had properties that differed dramatically from those of ‘‘classical’’ central systems6—by refusing to align with the prototypical bimodal characterization of being either an excitatory or an inhibitory neurotransmitter. The current status of these issues is comprehensively addressed in this handbook (see Gerfen, Chapter 2.1; Sesack Chapter 2.2; Caron and Gainetdinov, Chapter 3.2; Malenka et al., Chapter 7.2; and Surmeier et al., Chapter 7.3, this volume). Thus, my overview of the progress here will be highly selective in order to address two issues: (1) When do DA neurons fire? and (2) What does DA do to the postsynaptic targets of DA neurons?

1: OVERVIEW

WHEN DO DA NEURONS FIRE? The function of a DA neuron, like that of any other neuron, is established by its innate electrophysiological properties and by their synaptic input, which determines the activity of the system. Dopamine neurons in rats were initially localized post hoc by marking the recording sites using any of several methods and examining the recording sites cytologically after the experiment (see36 and Grace, Chapter 10.6, this volume). However, they were soon found to have a unique and reliable electrophysiological signal that allowed research to progress more rapidly, first in anesthetized animals and somewhat later in awake, behaving animals.37 Their signature activity consisted of polyphasic action potentials, initially positive or negative waveforms, followed by a prolonged positive component with a relatively long duration (1.8–3.6 ms) and an irregular firing at low baseline frequencies (0.5–8.5 spikes/s). In several experiments, it was possible to confirm the changes in activity in striatum and nucleus accumbens by microdialysis of DA.38 These early studies confirmed that DA neuron firing was slow in its basal state but shifted to burst firing when significant amounts of extracellular DA were detected. When studies of DA neuron function were performed in the nonhuman primate model, the activity of the neurons was also characterized by their electrophysiological signature, which was similar to that previously observed in the rat.39 This transition has proven to be very fruitful in that the greater cognitive skills of the primate permitted much more complex behavioral paradigms beyond drug self-administration and revealed that DA neuronal firing was not simply a signal that a reward to a behavioral response had been generated, but instead suggested that the fluctuating output of the primate DA neuron apparently signals changes or errors in the predictions of future salient and rewarding events. This has led to the emergence of quantitative theories of adaptive optimizing control.40,41 An interesting question is how DA, released from axon collaterals and dendrites, regulates the activity of the DA neuron. Using whole neuron recordings in slices from mouse substantia nigra and the ventrotegmental area, Beckstead and Williams42 observed an inhibitory postsynaptic current (IPSC) that was elicited by localized electrical stimulation of nearby DA neurons This IPSC was tetrodotoxin sensitive, calcium dependent, and blocked by a D2 receptor antagonist. Inhibition of monoamine transporters prolonged the IPSC, indicating that the time course of DA neurotransmission is tightly regulated by reuptake. Changing the stimulus intensity altered the amplitude but not the time course of the

5

IPSC, whose onset was faster than could be reproduced with iontophoresis. The results indicate a rapid rise in local DA concentration at the D2 receptors, suggesting that the DA that is released by a train of action potentials acts in a localized somatodendritic area, observations the authors conclude are incompatible with volume transmission.

WHAT DOES DA DO TO THE POSTSYNAPTIC TARGETS OF DA NEURONS? My singular adventure into the terminal fields of the nigra neurons focused on a question frequently raised at the beginning of cellular neuropharmacology, namely, ‘‘What does DA do?’’ In the parlance of the times, this question really meant ‘‘Is DA an excitatory or an inhibitory transmitter?’’ since those were the only possibilities then thought to exist. The caudate nucleus had been shown to be the region of the brain that had the highest content of DA, and we elected to probe for its effects on the spontaneous activity of caudate neurons in the cat because the caudate nucleus is impossible to miss with stereotaxic micromanipulators. The challenge of these experiments became immediately obvious when we could find virtually no continuously active spontaneous neurons in intact but anesthetized cats, and so before we could ask what DA did, we needed more active neurons on which to test it. In order to avoid the effects of general anesthetics, we resorted to a surgical method of forebrain isolation, which, unbeknown to us then, also severed the nigrostriatal pathway and greatly improved spontaneous activity. With these slowly discharging striatal units, we observed that acetylcholine or glutamate enhanced activity, while DA not only slowed the discharge, it did so for a prolonged period after its iontophoretic application, during which time excitatory responses to acetylcholine gradually returned.1 This was our starting point in confirming that DA did indeed act differently than the effects of other transmitters known at that time. As this handbook makes abundantly clear, the neuropharmacology of DA cannot be considered meaningful without characterization of the receptor subtypes under study, which sets the stage for the intracellular transductive pathways through which the transmitter is acting, and the range of pharmacological tools with which to simulate or antagonize the effects of DA (see Sibley et al., Chapter 3.1; and Fisone, Chapter 3.3, this volume). However, as with the conceptualizations of when DA neurons fire and what that firing connotes to the individual whose behavior is under observation, major steps

6

OVERVIEW

forward in elucidating the effects of DA on its target neurons in the nonhuman primate prefrontal cortex (PFC) also resulted from studies recording neuron activity and behavior. The meticulous experimental work of Sawaguchi and Goldman-Rakic,43,44 first demonstrated with a delayed oculomotor response task, showed that when the D1 receptor antagonists SCH23390 and SCH39166 were injected into the PFC, rhesus monkeys increased their errors and increased performance latency in a task that required memoryguided saccades, in a dose-dependent manner that was proportional to the duration of the delay period, but the D1 antagonists had no effect on performance in a control task requiring visually guided saccades. Based on a variety of other evidence including aging, disease, prior drug treatments, and experience, Sawaguchi and Goldman-Rakic proposed that when cortical DA was reduced below normal levels of function, a D1 agonist would improve function, while when cortical DA function was excessive, as hypothesized in schizophrenia, a D1 antagonist would improve function. When the Goldman-Rakic group turned their efforts to the functional effects of the D2 receptor in PFC using similar methods with behavioral tasks employing a sequence of phasic and tonic activations linked to a train of sensory, mnemonic, and response-related events, they observed that the DA D2 receptor selectively modulated the neural activities associated with memoryguided saccades in oculomotor delayed-response tasks but had little or no effect on the persistent mnemonicrelated activity regulated by the D1 receptors45 (also see Arnsten et al., Chapter 5.3, this volume). Still, for some observers, the question of what DA does remains, and for those for whom this query remains unanswered, recent observations on neuronal actions in the mouse caudate nucleus should help settle the issue. Not incidentally, these recent experiments were empowered by observations derived from neuronal localizations through stains. The vast majority of neurons in the striatum are medium spiny neurons, so named for the dendritic spines that characterize their morphology. Their organization was termed a complex mosaic organization (see Gerfen, Chapter 2.1, this volume). Through combinations of tract tracing and experimental lesions, it was recognized that the output circuitry of the caudate consists of two principal pathways: a striatonigral projection to the substantia nigra and the entopeduncular nucleus (referred to as the external pathway) and a striatopallidal projection to the globus pallidus (referred to as the internal pathway). Although all medium spiny neurons contain GABA, their associated neuropeptides and DA receptor subtype differ, depending on which output pathway they are in:

by immunocytochemistry and in situ hybridization, the striatonigral neurons express the neuropeptides substance P, and dynorphin, as well as the D1 receptor, while the striatopallidal neurons express proenkephalin and the D2 receptor. By developing transgenic mice in which the expression of D1 or D2 receptors was molecularly reported by the coexpression of green fluorescent protein, it was possible for Surmeier and colleagues46 to investigate two forms of striatal synaptic plasticity: long-term potentiation (LTP) and long-term depression (LTD). By controlling the sequences of presynaptic and postsynaptic activity in striatal slices with small microelectrodes that stimulated either afferent fibers close to a neuron or the neuron itself, when presynaptic activity precedes postsynaptic activity, LTP is produced, and when the sequence is reversed, LTD is produced. D2-expressing neurons were capable of reacting with either LTP or LTD according to the stimulation sequences, and here D2, but not D1, antagonists could block both potentiations. In D1-expressing spiny neurons, an LTP mediated by N-methyl-daspartic acid (NMDA) receptors was observed, but LTD was difficult to produce unless the D1 receptors were blocked with SCH23390; this LTD was, in turn, antagonized by blockade of the metabotropic Glutamate (GLU) type 5 receptor (mGluR5). These observations indicate that DA ‘‘is critical for the induction of the plasticity’’ acting in concert with GLU, adenosine, and activity in the external world. In conditions in which there are few if any behaviorally interesting stimuli, DA neurons fire slowly to keep highaffinity D2 receptors activated but not low-affinity D1 receptors; the latter are engaged when DA neurons fire in bursts to raise DA levels. Thus, the direction of the plasticity shaped by the same transmitter under different conditions will have distinct but consistent effects. These sorts of modulatory effects help establish in my mind, if not for others, the dynamic synaptic vocabulary for monoamine neurons that I have long envisioned.47

REFERENCES 1. Bloom FE, Costa E, Salmoiraghi GC. Anesthesia and the responsiveness of individual neurons of the caudate nucleus of the cat to acetylcholine, norepinephrine and dopamine administered by microelectrophoresis. J Pharmacol Exp Ther. 1965;150:244–252. 2. Bloom FE, Algeri S, Groppetti A, Revuelta A, Costa E. Lesions of central norepinephrine terminals with 6-OH-dopamine: biochemistry and fine structure. Science. 1969;166:1284–1286. 3. Bloom FE, Costa E. The effects of drugs on serotonergic nerve terminals. Adv Cytopharmacol. 1971;1:379–395.

1: OVERVIEW 4. Bloom FE, Costa E, Salmoiraghi GC. Analysis of individual rabbit olfactory bulb neuron responses to the microelectrophoresis of acetylcholine, norepinephrine and serotonin synergists and antagonists. J Pharmacol Exp Ther. 1965;146:16–23. 5. Bloom FE, Von B, Oliver AP, Costa E, Salmoiraghi GC. Microelectrophoretic studies of adrenergic mechanisms of rabbit olfactory neurons. Life Sci. 1964;3:131–136. 6. Moore RY, Bloom FE. Central catecholamine neuron systems: anatomy and physiology of the dopamine systems. Annu Rev Neurosci. 1978;1:129–169. 7. Dahlstro¨m A, Fuxe K. A method for the demonstration of monoamine-containing nerve fibres in the central nervous system. Acta Physiol Scand. 1964;60:293–294. 8. Dahlstro¨m A, Fuxe K. Evidence for the existence of monoamine-containing neurons in the central nervous system. I. Demonstration of monoamines in the cell bodies of brain stem neurons. Acta Physiol Scand Suppl. 1964;232(suppl):231–255. 9. Ungerstedt U. On the anatomy, pharmacology, and function of the nigro-striatal dopamine system. Thesis: Stockholm: The Karolinska Institute; 1971:122. 10. Fuxe K, Hoekfelt T, Nilsson O. Observations on the cellular localization of dopamine in the caudate nucleus of the rat. Z Zellforsch Mikrosk Anat. 1964;63:701–706. 11. Axelsson S, Bjorklund A, Falck B, Lindvall O, Svensson LA. Glyoxylic acid condensation: a new fluorescence method for the histochemical demonstration of biogenic monoamines. Acta Physiol Scand. 1973;87:57–62. 12. Lindvall O, Bjorklund A, Hokfelt T, Ljungdahl A. Application of the glyoxylic acid method to vibratome sections for the improved visualization of central catecholamine neurons. Histochemie. 1973;35:31–38. 13. Ho¨kfelt T, Johansson O, Fuxe K, Goldstein M, Park D. Immunohistochemical studies on the localization and distribution of monoamine neuron systems in the rat brain. I. Tyrosine hydroxylase in the mes- and diencephalon. Med Biol. 1976;54:427–453. 14. Ho¨kfelt T, Johansson O, Fuxe K, Goldstein M, Park D. Immunohistochemical studies on the localization and distribution of monoamine neuron systems in the rat brain II. Tyrosine hydroxylase in the telencephalon. Med Biol. 1977;55:21–40. 15. Pickel VM, Joh TH, Reis DJ. Monoamine-synthesizing enzymes in central dopaminergic, noradrenergic and serotonergic neurons. Immunocytochemical localization by light and electron microscopy. J Histochem Cytochem. 1976;24:792–306. 16. Smiley JF, Goldman-Rakic PS. Silver-enhanced diaminobenzidine-sulfide (SEDS): a technique for high-resolution immunoelectron microscopy demonstrated with monoamine immunoreactivity in monkey cerebral cortex and caudate. J Histochem Cytochem. 1993;41:1393–1404. 17. Smiley JF, Goldman-Rakic PS. Heterogeneous targets of dopamine synapses in monkey prefrontal cortex demonstrated by serial section electron microscopy: a laminar analysis using the silverenhanced diaminobenzidine sulfide (SEDS) immunolabeling technique. Cereb Cortex. 1993;3:223–238. 18. Dowling JE. Retinal neuromodulation: the role of dopamine. Vis Neurosci. 1991;7:87–97. 19. Ljungdahl A, Hokfelt T, Halasz N, Johansson O, Goldstein M. Olfactory bulb dopamine neurons – the A15 catecholamine cell group. Acta Physiol Scand Suppl. 1977;452:31–34. 20. Halasz N, Hokfelt T, Ljungdahl A, Johansson O, Goldstein M. Dopamine neurons in the olfactory bulb. Adv Biochem Psychopharmacol. 1977;16:169–177.

7

21. Cransac H, Peyrin L, Cottet-Emard JM, Farhat F, Pequignot JM, Reber A. Aging effects on monoamines in rat medial vestibular and cochlear nuclei. Hear Res. 1996;100:150–156. 22. Darrow KN, Simons EJ, Dodds L, Liberman MC. Dopaminergic innervation of the mouse inner ear: evidence for a separate cytochemical group of cochlear efferent fibers. J Comp Neuro.l 2006;498:403–414. 23. Pivonello R, Ferone D, Lombardi G, Colao A, Lamberts SW, Hofland LJ. Novel insights in dopamine receptor physiology. Eur J Endocrinol. 2007;156(suppl 1):S13-S21. 24. Berger B, Thierry AM, Tassin JP, Moyne MA. Dopaminergic innervation of the rat prefrontal cortex: a fluorescence histochemical study. Brain Res. 1976;106:133–145. 25. Berger B, Tassin JP, Blanc G, Moyne MA, Thierry AM. Histochemical confirmation for dopaminergic innervation of the rat cerebral cortex after destruction of the noradrenergic ascending pathways. Brain Res. 1974;81:332–337. 26. Thierry AM, Blanc G, Sobel A, Stinus L, Glowinski J. Dopaminergic terminals in the rat cortex. Science. 1973;182:498–500. 27. Lewis DA. The catecholaminergic innervation of primate prefrontal cortex. J Neural Transm Suppl. 1992;36:179–200. 28. Lewis DA, Sesack SR. Dopamine systems in the primate brain. In: Bloom FE, Hokfelt T, eds. The Primate Nervous System, Part I. Amsterdam: Elsevier; 1997:187–216. 29. Beckstead MJ, Grandy DK, Wickman K, Williams JT. Vesicular dopamine release elicits an inhibitory postsynaptic current in midbrain dopamine neurons. Neuron. 2004;42:939–946. 30. Chau DT, Roth RM, Green AI. The neural circuitry of reward and its relevance to psychiatric disorders. Curr Psychiatry Rep. 2004;6: 391–399. 31. Tam SY, Roth RH. Mesoprefrontal dopaminergic neurons: can tyrosine availability influence their functions? Biochem Pharmacol. 1997;53:441–453. 32. Horger BA, Roth RH. The role of mesoprefrontal dopamine neurons in stress. Crit Rev Neurobiol. 1996;10:395–418. 33. Lewis DA, Gonzalez-Burgos G. Pathophysiologically based treatment interventions in schizophrenia. Nat Med. 2006;12:1016–1022. 34. Goldman-Rakic PS. The physiological approach: functional architecture of working memory and disordered cognition in schizophrenia. Biol Psychiatry. 1999;46:650–661. 35. Goldman-Rakic PS, Castner SA, Svensson TH, Siever LJ, Williams GV. Targeting the dopamine D1 receptor in schizophrenia: insights for cognitive dysfunction. Psychopharmacology (Berl). 2004;174:3–16. 36. Grace AA. Dopamine. In: Davis KL, Charney D, Coyle JT, Nemeroff C, eds. Neuropsychopharmacology—The Fifth Generation of Progress. Philadelphia, PA: Lippincott-Williams & Wilkins; 2002:119–132. 37. Diana M, Garcia-Munoz M, Richards J, Freed CR. Electrophysiological analysis of dopamine cells from the substantia nigra pars compacta of circling rats. Exp Brain Res. 1989;74:625–630. 38. Hurd YL, Weiss F, Koob GF, And NE, Ungerstedt U. Cocaine reinforcement and extracellular dopamine overflow in rat nucleus accumbens: an in vivo microdialysis study. Brain Res. 1989;498:199–203. 39. Schultz W, Romo R. Dopamine neurons of the monkey midbrain: contingencies of responses to stimuli eliciting immediate behavioral reactions. J Neurophysiol. 1990;63:607–624. 40. Kobayashi S, Schultz W. Influence of reward delays on responses of dopamine neurons. J Neurosci. 2008;28:7837–7846.

8

OVERVIEW

41. Fiorillo CD, Newsome WT, Schultz W. The temporal precision of reward prediction in dopamine neurons. Nat Neurosci. 2008;11:966–973. 42. Beckstead MJ, Williams JT. Long-term depression of a dopamine IPSC. J Neurosci. 2007;27:2074–2080. 43. Sawaguchi T, Goldman-Rakic PS. D1 dopamine receptors in prefrontal cortex: involvement in working memory. Science. 1991;251:947–950. 44. Sawaguchi T, Goldman-Rakic PS. The role of D1-dopamine receptor in working memory: local injections of dopamine antagonists into the prefrontal cortex of rhesus monkeys

performing an oculomotor delayed-response task. J Neurophysiol. 1994;71:515–528. 45. Wang M, Vijayraghavan S, Goldman-Rakic PS. Selective D2 receptor actions on the functional circuitry of working memory. Science. 2004;303:853–856. 46. Shen W, Flajolet M, Greengard P, Surmeier DJ. Dichotomous dopaminergic control of striatal synaptic plasticity. Science. 2008;321:848–851. 47. Bloom FE. Dynamic synaptic communication: finding the vocabulary. Brain Res. 1973;62:299–305.

2 Neuroanatomy

This page intentionally left blank

2.1

Functional Neuroanatomy of Dopamine in the Striatum CHARLES R. GERF EN

From the perspective of a neuroanatomist, a major significance of the discovery of dopamine as a neurotransmitter was the subsequent development of the fluorescent catecholamine histochemical methods, which established the ability to visualize neuroanatomical circuits on the basis of their neurotransmitter.1,2 In the early 1960s, Swedish researchers had the insight to exploit the ability to convert catecholamines into fluorescent molecules by a condensation reaction with formaldehyde to develop a histochemical method that revealed catecholamine-containing neurons and their axons.3,4 Axonal tracing studies identified additional details of the organization of the nigrostriatal dopamine system and other neuroanatomical circuits of the basal ganglia. Further, the introduction of immunohistochemical methods in the late 1970s and molecular techniques employing in situ hybridization histochemical localization of messenger RNAs in the late 1980s generalized the ability to map the neuroanatomical organization of neurochemically characterized circuits. These latter techniques have provided considerable understanding of the functional circuits of the basal ganglia, built upon the pioneering work using the catecholamine histofluorescence techniques.

single continuous system whose axonal projections provide a dense input to all parts of the striatum, including the nucleus accumbens. The projection of the mesostriatal dopamine system is topographically organized such that medially located neurons project ventrally to the nucleus accumbens and ventral striatum, and projections to the dorsal striatum originate from more laterally positioned neurons in the substantia nigra pars compacta and the retrorubral area. While the most prominent projection of the mesostriatal dopamine neurons is to the striatum, some neurons in this complex also provide inputs to other forebrain areas including the cerebral cortex. Dopamine axons are densely and rather homogeneously distributed in the striatum. In turn, the neurons of the striatum itself are distributed homogeneously, without any obvious cytoarchitectural features. There is, for example, no clear cytoarchitectural feature that separates the nucleus accumbens from the ventral parts of the striatum. However, this homogeneity in both the dopamine input system and the distribution of striatal neurons masks a number of underlying neuroanatomical circuits that are key to understanding the functional organization of the basal ganglia.

THE MESOSTRIATAL DOPAMINE SYSTEM

STRIATAL PATCH-MATRIX COMPARTMENTS

The most prominent of the dopamine systems revealed by the catecholamine histochemical techniques is the group of dopamine neurons in the midbrain, which provide a dense axonal projection to the striatum and the nucleus accumbens.3,4 Three cell groups were originally identified based on their location: the A10, A9, and A8 dopamine cell groups located respectively in the ventral tegmental area, substantia nigra pars compacta, and retrorubral area. For the most part, there are no clear boundaries between these groups; they form a

Indications that there are compartments within the striatum came from studies that observed islands or patches of dopamine innervation distributed within the neuropil of the striatum during early postnatal development that give way to a homogeneous distribution as development progresses.5 A number of neurochemical markers were found to coincide with these patches, including staining for acetylcholinesterase6 and opiate receptor binding.7 The striatal neurons that are the target of this early dopamine input also develop first 11

12

NEUROANATOMY

within the striatum, with later-born striatal neurons filling in the surrounding matrix regions of the striatum.8 These two developmental compartments, the early-developing patches or islands and the laterdeveloping matrix, give rise to the adult patch (or striosome)6 and matrix compartments of the striatum. As the adult striatum appears homogeneous, neurochemical markers are required to reveal them; notably, among others, calbindin marks the matrix compartment9 and mu opiate receptors mark the patch compartment.7 Distinct subsets of dopamine neurons differentially target the striatal patch and matrix compartments10,11 (Fig. 2.2.1). While the mesostriatal dopamine neurons in the ventral tegmental area, susbstantia nigra pars compacta, and retrorubral area appear as a continuous grouping of neurons, within the substantia nigra pars compacta there are two parts that form a dorsal and a ventral tier of neurons. Dorsal tier pars compacta dopamine neurons are continuous with ventral tegmental dopamine neurons and extend their dendrites in the plane of the pars compacta medial and laterally. Ventral tier dopamine neurons are organized in two parts. One part is immediately ventral to the dorsal tier in the pars compacta and extend its dendrites ventrally into the substantia nigra pars reticulata. The second part consists of dopamine neurons that are grouped within the substantia nigra pars reticulata itself. Axonal tracing studies demonstrated that dopamine neuron projections from the ventral tegmental area, dorsal tier of the substantia nigra pars compacta, and retrorubral area provide input to the striatal matrix compartment, whereas projections from the two groups of ventral tier dopamine neurons of the substantia nigra provide input to the striatal patch compartment.10 Moreover, matrix projecting neurons coexpress the calcium binding protein, calbindin, which provides a neurochemical marker for these dopamine neurons.9,11 To confirm this organization, we took advantage of the differential development of the patch- and matrix-directed dopamine systems. Injecting the neurotoxin 6-hydroxydopamine into the striatum on the day of birth resulted in the selective degeneration of the ventral tier dopamine neurons and the dopamine input to the patch compartment.11 As adults, the calbindin-expressing dopamine neurons in the ventral tegmental area, dorsal tier of the pars compacta, and retrorubral area survived, as did the dopamine input to the striatal matrix compartment. These studies in the rat demonstrate distinct sets of mesostriatal dopamine neurons that differerentially target the striatal patch and matrix compartments, which has also been demonstrated in the primate.9,12 A recent study by Matsuda et al.13 adds important details concerning the compartmental organization of the nigrostriatal

dopamine system. Using a method that labels the full axonal arborization of single neurons, these authors show that individual dopamine neurons in both the dorsal and ventral tiers distribute axons to both striatal patch and matrix compartments, although each neuron’s arborization tend to favor one or the other.

INPUT-OUTPUT ORGANIZATION OF THE STRIATAL PATCH AND MATRIX COMPARTMENTS In addition to dopaminergic inputs, other input and output connections of the striatum are organized relative to the patch-matrix compartments (Fig. 2.1.2). The major neuron type in the striatum is the medium spiny neuron, which constitutes up to 90% of the striatal neuron population. These neurons provide the axonal output of the striatum, with different populations of these neurons targeting different basal ganglia nuclei. Although the distribution of these neurons is homogeneous and does not reveal striatal compatments, the dendrites of the neurons in patch and matrix compartments remain confined within their respective compartments.14,15 Axonal tracing studies demonstrate that projections from the striatal patch compartment provide input directed principally to the ventral tier dopamine neurons in the substantia nigra, whereas the striatal matrix neurons project to the globus pallidus, entopeduncular nucleus, and substantia nigra pars reticulate.14 Thus, the striatal output of the patch compartment is directed principally at the same ventral tier dopamine neurons that provide input to this compartment. In this regard, the recent finding that ventral tier dopamine neurons provide dopamine input to both patch and matrix compartments13 is important. This finding suggests that the striatal patch output is not part of a closed loop with the dopamine neurons providing patch input, but rather affects dopamine feedback to both compartments. The target of the output of striatal matrix neurons is directed to components of the basal ganglia that provide the output of this system. In particular, the entopeduncular nucleus and substantia nigra pars reticulata are composed of GABAergic neurons, which project to the thalamus and superior colliculus and other midbrain systems connected with motor control. Thus, the output of neurons in the striatal patch and matrix respectively target dopamine feedback to the striatum and basal ganglia output systems. Dopamine input to striatal medium spiny neurons is directed principally to dendritic shafts and spine necks and likely functions to modulate excitatory input that is directed to the dendritic spines. There are two main

FIGURE 2.1.1. The organization of the nigrostriatal dopamine (DA) pathway from the midbrain to the striatum (sagittal diagram at upper right) is diagrammed to show the organization of this system to the striatal patch and matrix compartments. Coronal sections at three levels through the striatum (A) are depicted to show the innervation of the patch and matrix compartments from different subsets of midbrain DA neurons from three levels (B, C, D). Neurons providing inputs to the striatal matrix compartment (white in B, C, D) are located in the ventral tegmental area (VTA, A10 DA cell group), in the dorsal tier of the substantia nigra pars compacta (in B,C: SNCD, A9), and in the retrorubral area (in D: RRF, A8 DA cell group). Neurons providing input to the striatal patch compartment are located in the ventral tier of the substantia nigra pars compacta (in B, C, D: DA neurons in dark gray areas) and project from A9 DA cells located in the substantia nigra pars reticulata (in C and D). There is a general topography in that medially located cells project to the ventral striatum and laterally located cells project to the dorsal striatum. Neurons at each rostral-caudal level in the midbrain project rather extensively throughout the rostral-caudal extent of the striatum. (See Color Plate 2.1.1.)

13

14

NEUROANATOMY

FIGURE 2.1.2. Organization of the striatal patch-matrix compartments provides parallel pathways from the cerebral cortex through the striatum that provide differential input to the dopamine and GABA neurons in the substantia nigra. Deep layer 5 corticostriatal neurons provide selective inputs to the striatal patch compartment, whose neurons provide inputs targeting dopamine neurons in the substantia nigra pars compacta. Superficial layer 5 corticostriatal neurons provide inputs to the striatal matrix compartment, whose neurons project to the substantia nigra pars reticulata, which contains the GABAergic output neurons of the basal ganglia. This organization arises from most neocortical areas, although there is a gradient such that those areas closer to the allocortex provide greater input to the patch compartment, whereas primary sensorimotor areas provide greater input to the matrix compartment. (See Color Plate 2.1.2.)

sources of glutamatergic excitatory input, the cerebral cortex and thalamus, and some aspects of each of these are organized relative to patch and matrix compartments. For the thalamus, parts of the intralaminar thalamic nuclei differentially target the patch matrix compartments, with projections of the parafascicular and centromedian nuclei directed to the matrix and projections of the paraventricular nucleus directed to the patch compartment.7,16,17 Corticostriatal projections arise from pyramidal neurons in layer 5. Initial studies examining the compartmental targets of corticostriatal projections suggested that different cortical areas projected

selectively to either patch or matrix. Limbic cortical areas were shown to provide input to the patch compartment, whereas neocortical somatosensory and motor cortical area projections targeted the matrix.18,19 However, more detailed analysis of corticostriatal projections demonstrated that most cortical areas provided inputs to both compartments, but that neurons in different sublayers of layer 5 differentially project to the patch and matrix compartments.20 For each specific cortical area, neurons with patch-directed inputs are located in deep layer 5, whereas those with matrix-directed inputs are located in superficial layer 5. Corticostriatal projections are topographically

2.1: FUNCTIONAL NEUROANATOMY OF DOPAMINE IN THE STRIATUM

organized such that motor cortical areas project to the dorsolateral striatum and prelimbic and infralimbic areas project to the medial and ventral striatum. Importantly, for each cortical area, both patch- and matrix-directed projections target the topographic region within the striatum such that from a given cortical area, its projections to the matrix surround the patches to which it also projects. While this pattern of organization of the corticostriatal projections is apparent in most cortical areas, the relative contribution of inputs to the patch and matrix compartments varies among cortical areas. Neocortical areas, such as motor, supplementary motor, and somatosensory cortices, provide greater inputs to the matrix compartment, whereas allocortical and periallocortical areas such as the prelimbic and infralimbic cortical areas provide greater inputs to the patch compartment. This transition of a predominance of patch-directed inputs from limbic-related cortical areas to matrixdirected inputs from neocortical areas is likely responsible for the earlier findings suggesting that different cortical areas provide inputs only to one compartment. The major significance of the organization of corticostriatal projections is that the striatal patch and matrix compartments are related to the laminar organization of the cerebral cortex rather than to tangential or columnar features of its organization.20

D1AND D2 DOPAMINE RECEPTORS IN DIRECT AND INDIRECT STRIATAL PROJECTIONS A major discovery concerning the function of dopamine in the basal ganglia was the demonstration that D1 and D2 dopamine receptors are segregated in the direct and indirect striatal projection neurons21 (Fig. 2.1.3). Striatal medium spiny neurons, which constitute up to 90% of the neuron population of the striatum and nucleus accumbens, are composed of two major subtypes based on their axonal projections. One subtype projects axons through the globus pallidus, making some contacts, but extends axons to terminate in the internal segment of the globus pallidus (or entopeduncular nucleus) and substantia nigra. These nuclei constitute the major output system of the basal ganglia such that the striatal neurons that project to them directly are considered to provide the ‘‘direct’’ striatal projection pathway. The other subtype of striatal projection neuron extends its axon only to the globus pallidus. Neurons in this nucleus provide inputs to the internal segment of the globus pallidus and substantia nigra and to the subthalamic nucleus, which in turn projects to these basal ganglia output nuclei. Thus, striatal neurons

15

that project only to the globus pallidus are connected through multiple synaptic connections to the output of the basal ganglia and are considered to provide the ‘‘indirect’’ striatal projection pathway. Neurons giving rise to the direct and indirect pathway are approximately equal in number and are intermingled with one another in both the patch and matrix compartments.22 The functional significance of the striatal direct and indirect pathways was established by the observation that following dopamine depletion in the striatum, there are differential changes in GABA receptor binding in the globus pallidus and substantia nigra23 and in the expression of peptides expressed by striatal direct and indirect pathway neurons.24 These findings led to the hallmark theory that clinical movement disorders such as Parkinson’s disease result from an imbalance in the output activity of the striatal direct and indirect pathways.25,26 This theory suggested that akinesia, which characterizes Parkinson’s disease, is a consequence of increased functional activity in the indirect striatal pathway. The underlying mechanism responsible for dopamine-mediated differential changes in the functional activity of the striatal direct and indirect pathways was shown to be that D1- and D2-dopamine receptors are respectively segregated in the neurons giving rise to these projections.21 Two lines of evidence were provided in this study. The first line of evidence was provided by neuroanatomical studies. In situ hybridization histochemical localization of the mRNAs encoding D1dopamine receptors demonstrated the selective expression of these receptors in neurons that project to the substantia nigra and coexpress the peptides dynorphin and substance P, markers of the direct striatal pathway. On the other hand, D2 mRNA was shown to be expressed selectively in neurons that project to the globus pallidus and coexpress the peptide enkephalin, a marker of indirect striatal pathway neurons. The second line of evidence was provided by functional studies. Following dopamine depletion of the nigrostriatal pathway, enkephalin expression increases in indirect pathway neurons, whereas substance P and dynorphin expression decreases in direct pathway neurons. These dopamine lesion–induced changes in gene expression were demonstrated to be selectively reversed in indirect pathway neurons with D2 receptor agonist treatment and in direct pathway neurons with D1 receptor agonist treatment. This finding was somewhat controversial, as some investigators maintained that D1 and D2 dopamine receptors are coexpressed in most striatal medium spiny neurons.23 However, the segregation of D1 and D2 dopamine receptors in direct and indirect striatal pathway neurons has been confirmed by numerous

16

NEUROANATOMY

FIGURE 2.1.3. Circuitry involved in Parkinson’s disease. Upper diagram: Imbalances in the function of direct and indirect pathways of the basal ganglia in Parkinson’s disease, shown in a sagittal brain section of the mouse. The cerebral cortex and thalamus provide excitatory inputs to the striatum, the main input nucleus of the basal ganglia. The output of the basal ganglia originates from the medial globus pallidus (GPm) and the substantia nigra pars reticulata (SNr) and is directed primarily to thalamic nuclei, which project to frontal areas of the cerebral cortex. The direct pathway originates from striatal projection neurons whose axons extend directly to the GPm and SNr output nuclei. The indirect pathway originates from striatopallidal neurons whose axons terminate within the globus pallidus (GP). Neurons in the GP, in turn, project to the subthalamic nucleus (STN), which projects to the GPm and SNr. Thus, striatopallidal neurons are connected indirectly, through the GP and STN, with the output of the basal ganglia. Lower images: D1 and D2 dopamine neurons are segregated to direct- and indirect-pathway neurons, respectively. Sagittal sections from BAC transgenic mice in which these receptors are labeled with enhanced green fluorescence protein (EGFP) show labeling of the neuron cell bodies in the striatum as well as their axonal projections. D2-BAC transgenic mice show labeling of the indirectpathway neurons (these axon projections terminate in the GP), whereas D1-BAC mice show labeling of the direct pathway, as seen by labeling of axon terminals in the GPm and SNr.30,31 (See Color Plate 2.1.3.)

studies such that there is now a consensus in the field upholding the original finding.27–31 The demonstration of segregation of D1 and D2 dopamine receptors, respectively, in direct and indirect pathway neurons21 provided the basis for understanding the functional changes in movement disorders

such as Parkinson’s disease.25,26 The central tenet of the theory of movement disorders is that they result from imbalanced activity in the direct and indirect striatal pathways. In Parkinson’s disease, which is marked by akinesia, the theory suggests that there is increased activity in the indirect pathway. Neurons of this

2.1: FUNCTIONAL NEUROANATOMY OF DOPAMINE IN THE STRIATUM

pathway express the D2 dopamine receptor, which is coupled to the inhibitory G protein, Gi. In the normal animal, dopamine binding to the D2 receptors provides an inhibitory function. On the other hand, the D1 receptor expressed on direct pathway neurons is coupled to stimulatory G proteins, Gs and Golf. Consequently, in Parkinson’s disease, the loss of dopamine input to the striatum has opposite effects on the direct and indirect pathways, with increased function in the indirect pathway and decreased function in the direct pathway. Surgical therapies developed to reverse this imbalance by interfering with altered function in the indirect pathway proven to have considerable clinical benefit.32,33

L-DOPA-INDUCED DYSKINESIA IN PARKINSON’S DISEASE Treatment of Parkinson’s disease with L-DOPA34 was a direct result of the discovery of dopamine and its depletion in the disease and remains the primary therapeutic treatment. While it is a very effective therapy, long-term treatment invariably leads to the development of dyskinesias.35 We have proposed that LDOPA-induced dyskinesia in the treatment of Parkinson’s disease results from an aberrant switch in the linkage of the D1 dopamine receptor to signal transduction systems that activate the protein kinase, extracellular signal-regulated protein kinase (ERK1/2).36 As discussed, dopamine depletion of the striatum results in opposite effects on the function of D2-indirect and D1-direct pathway striatal neurons evidenced by changes in gene expression.21 While either L-DOPA or selective D2 and D1 receptor agonist treatments reverse some of the gene expression changes, the response of D1-receptor-bearing direct pathway neurons is supersensitive to these treatments, as demonstrated by the induction of a large number of so-called immediate early genes.37 The supersensitive response of striatal neurons following lesioning of the nigrostriatal dopamine system was first described by Ungerstedt,38 who observed that animals with unilateral lesions exhibited a robust rotation contralateral to the lesioned side in response to direct and indirect dopamine agonist treatment. This experimental paradigm remains the standard animal model for the study of Parkinson’s disease. The reasonable explanation of why these animals display contralateral rotation following dopamine receptor agonist treatment was that striatal neurons compensated for the loss of dopamine by increasing their expression of dopamine receptors in order to increase their response to decreased levels of neurotransmitter. Thus, striatal neurons in the

17

lesioned striatum would produce a supersensitive response relative to the dopamine-intact striatum, which resulted in the behavioral rotation. However, our studies demonstrating the segregation of D1 and D2 dopamine receptors on direct and indirect striatal neurons provide a different model.21 Following dopamine lesioning, there is an increase in D2 dopamine receptor expression in indirect pathway neurons and a decrease in D1 receptor expression in direct pathway neurons. Rather than reflecting a compensatory response of striatal neurons to decreased dopamine input, these changes in receptor expression reflect the simple consequence of the loss of dopamine function on these neurons. Thus, in indirect pathway neurons, the absence of dopamine acting on D2 dopamine receptors, coupled to the inhibitory G protein, Gi, results in increased gene expression in these neurons, including the D2 dopamine receptor. On the other hand, in direct pathway neurons, the absence of dopamine acting on D1 dopamine receptors, coupled to the stimulatory G proteins, Gs and Golf, results in decreased gene expression, including the D1 dopamine receptor. In addition to the behavioral rotational response in the unilateral dopamine lesion paradigm, a cellular response was shown by the demonstration of the induction of immediate early genes (IEGs), such as c-fos, in striatal neurons in response to dopamine agonists.39,40 Significantly, the IEG response to L-DOPA or dopamine agonists such as apomorphine was found to occur exclusively in D1-expressing direct pathway striatal neurons. This IEG response provides a cellular measure of receptor supersensitivity. What is most interesting about this IEG response in D1 receptor–expressing neurons is that it occurs upon the first treatment with the dopamine receptor agonist, when the level of D1 receptor expression is decreased compared with that of neurons in the dopamine-intact striatum.37 This finding suggested that in the dopamine-lesioned striatum, the supersensitive response of D1 dopamine receptor– expressing neurons is not a consequence of increased D1 dopamine receptor expression, but rather is due to a change in the coupling of this receptor to signal transduction systems. Psychostimulants, such as cocaine and amphetamine, produce robust induction of IEGs in the normal dopamine-innervated striatum.41 This raises the question as to whether the D1 dopamine receptor–mediated supersensitive induction of IEGs in the dopamine-depleted striatum is due to an amplification of the normal D1 receptor coupling to signal transduction. However, psychostimulant striatal IEG induction differs in important ways from the D1 response in the dopamine-depleted striatum. First, whereas the psychostimulant response is

18

NEUROANATOMY

dependent on glutamate N-methyl-D-aspartate (NMDA) receptor activation,42 the D1-mediated IEG induction in the dopamine-depleted striatum occurs independently of NMDA receptor function.43 Second, repeated psychostimulant treatment produces an attenuated striatal IEG response44; the response in the dopamine-depleted striatum increases with extended dopamine-receptor agonist treatment.45 Using pharmacologic treatment paradigms to compare D1 receptor–mediated signaling in the dopamineintact and -lesioned striatum, activation of ERK1/2 was demonstrated to occur exclusively in the dopaminedepleted striatum.36 In this study, pharmacologic treatments with high doses of D1 receptor agonists, or combined D1 and D2 dopamine agonists, produced induction of IEGs in the dopamine-intact striatum at levels comparable to those produced in the dopaminedepleted striatum. However, activation of ERK1/2 occurred only in the dopamine-depleted striatum. In the dopamine-intact striatum, dopamine-agonist treatment activation of ERK1/2 was limited to the nucleus accumbens. This finding suggests that the depletion of dopamine in the striatum produces an aberrant coupling of the D1 receptor with activation of ERK1/2.

Psychostimulant treatments activate ERK1/2 in the nucleus accumbens and in a small percentage of neurons in the dorsal striatum, which was proposed to be dependent on dopamine- and cyclic adenosine monophosphate (cAMP)–regulated phosphoprotein, 32 kDa (DARPP32).46 However, in transgenic mice with either the D1 dopamine receptor or DARPP32 knocked out, psychostimulant activation of ERK1/2 occurs in the dorsal striatum but is reduced in the nucleus accumbens.37 Thus, psychostimulant activation of ERK1/2 in the dorsal striatum does not appear to involve the D1 dopamine receptor. Moreover, it is important to note that psychostimulant activation occurs in a small percentage of dorsal striatal neurons (approximately 10%); in the dopamine-depleted striatum, D1 dopamine receptor activation of ERK1/2 occurs in nearly all D1 receptor–expressing dorsal striatal neurons. On the other hand, in mice with the D1 dopamine receptor knocked out, L-DOPA treatment did not produce activation of ERK1/2 in the dopamine-depleted striatum. However, in mice with DARPP-32 knocked out, LDOPA treatment produced robust activation of ERK1/ 2 in the dopamine-depleted striatum similar to that in wild-type mice.47

Demonstration of distinct mechanisms of D1 dopamine receptor–mediated gene regulation in the dopamine (DA)-intact and-depleted striatum, using the full D1 agonist SKF81297 alone or combined with other drugs. (A–C) In situ hybridization histochemical localization of mRNA encoding c-fos 45 min after different drug combinations: A, SKF81297 (0.5 mg/kg); B, SKF81297 (2.0 mg/kg); C, SKF81297 (2.0 mg/kg) combined with the D2 DA receptor agonist (1 mg/kg) and scopolamine. The low dose of agonist alone (A) demonstrates the supersensitive response by the selective induction of c-fos in the DA-depleted striatum. Bilateral induction of c-fos IEG in both the DA-intact and -depleted striatum follows treatment with a high dose of the full D1 agonist alone (B) or in combination with other drugs (C). However, when animals receiving any of these treatments are killed at 15 min, p-ERK1/2- immunoreactive neurons are evident only in the DA-depleted striatum and not in the DA-intact striatum (data not shown). The treatment combining the full D1 agonist with both the D2 agonist and scopolamine produces a robust c-fos IEG response in both the DA-intact (D) and DA-depleted striatum (E). This treatment also results in persistent p-ERK1/2 (H) in the DA-depleted striatum but does not activate p-ERK1/2 (G) in neurons in the DA-intact striatum. These results demonstrate that, although D1-DA receptor-mediated induction of the IEG c-fos occurs in both the DA-intact and -depleted striatum, activation of ERK1/2 occurs only in the DA-depleted striatum.36 FIGURE 2.1.4.

2.1: FUNCTIONAL NEUROANATOMY OF DOPAMINE IN THE STRIATUM

19

FIGURE 2.1.5. Drd1a-agonist or L-DOPA activation of ERK1/2 in the dopamine (DA)-depleted striatum does not involve DARPP-32. Comparison of coronal brain sections at the level of the rostral striatum from wild-type and DARPP-32 knockout (KO) mice, with unilateral lesions of the nigrostriatal DA system and treated with a drd1a agonist (SKF81298, 5 mg/kg, 1 day) or L-DOPA (20 mg/kg with 12 mg/kg benserazide, 10 days). DARPP-32 immunoreactivity (IR) labels neurons in the striatum in wild-type mice, which are unlabeled in DARPP-32 KO mice. The unilateral lesion of the nigrostriatal DA pathway in these animals is shown by the absence of tyrosine hydroxylase (TH)-IR in the axonal terminals in the right striatum. Activation of ERK1/2 in response to either drd1a-agonist treatment (left-side figures) or L-DOPA treatment (right-side figures) is demonstrated by phospho-ERK1/2-IR throughout the DA-depleted striatum. High-power images from the dorsolateral striatum (inset boxes, 100 um wide) show few to no phospho-ERK1/2 IR neurons in the DA-intact striatum. In contrast, there are numerous phospho-ERK1/2 IR neurons in the DA-depleted striatum in both the wild- type and DARPP-32 KO animals.47

Together these studies suggest that following dopamine depletion, there is an aberrant coupling of the D1 dopamine receptor to activation of ERK1/2 that does not involve DARPP32. In addition, these studies point to distinct regional differences between the nucleus accumbens and dorsal striatum in the coupling of the D1 receptor with signal transduction systems. Based on these findings, we suggested that following dopamine depletion in the striatum, repeated activation of the aberrant coupling of the D1 dopamine receptor by LDOPA is responsible for the development of dyskinesias.36 Subsequent studies by a number of groups have demonstrated in Parkinson’s disease animal models strong correlations between L-DOPA induction of ERK1/2 and dyskinesias.48–50

REFERENCES 1. Carlsson A, Falck B, Hillarp NA. Cellular localization of brain monoamines. Acta Physiol Scand Suppl. 1962;56:1–28. 2. Dahlstroem A, Fuxe K. A method for the demonstration of monoamine-containing nerve fibers in the central nervous system. Acta Physiol Scand. 1964;60:293–294. 3. Anden NE, Carlsson A, Dahlstroem A, Fuxe K, Hillarp NA, Larsson K. Demonstration and mapping out of nigro-neostriatal dopamine neurons Life Sci. 1964;3:523–530. 4. Dahlstroem A, Fuxe K. Evidence for the existence of monoamine containing neurons in the central nervous system. I. Demonstration of monoamines in the cell bodies of brain stem neurons. Acta Physiol Scand. 1964(suppl);232:1–55. 5. Tennyson VM, Barrett RE, Cohen G, Coˆte´ L, Heikkila R, Mytilineou C. Correlation of anatomical and biochemical development of the rabbit neostriatum. Prog Brain Res. 1973;40:203–217.

20

NEUROANATOMY

6. Graybiel AM, Ragsdale CW Jr. Histochemically distinct compartments in the striatum of human, monkeys, and cat demonstrated by acetylthiocholinesterase staining. Proc Natl Acad Sci USA. 1978;75:5723–5726. 7. Herkenham M, Pert CB. Mosaic distribution of opiate receptors, parafascicular projections and acetylcholinesterase in rat striatum. Nature. 1981;291:415–418. 8. van der Kooy D, Fishell G. Neuronal birthdate underlies the development of striatal compartments. Brain Res. 1987;401:155–161. 9. Gerfen CR, Baimbridge KG, Miller JJ. The neostriatal mosaic: compartmental distribution of calcium-binding protein and parvalbumin in the basal ganglia of the rat and monkey. Proc Natl Acad Sci USA. 1985;82:8780–8784. 10. Gerfen CR, Herkenham M, Thibault J. The neostriatal mosaic: II. Patch- and matrix-directed mesostriatal dopaminergic and nondopaminergic systems. J Neurosci. 1987;7:3915–3934. 11. Gerfen CR, Baimbridge KG, Thibault J. The neostriatal mosaic: III. Biochemical and developmental dissociation of patch-matrix mesostriatal systems. J Neurosci. 1987;7:3935–3944. 12. Jimenez-Castellanos J, Graybiel AM. Subdivisions of the dopamine containing A8-A9-A10 complex identified by their differential mesostriatal innervation of striosomes and extrastriosomal matrix. Neuroscience. 1987;23:223–242. 13. Matsuda W, Furuta T, Nakamura KC, Hioki H, Fujiyama F, Arai R, Kaneko T. Single nigrostriatal dopaminergic neurons form widely spread and highly dense axonal arborizations in the neostriatum. J Neurosci. 2009;29:444–453. 14. Gerfen CR. The neostriatal mosaic. I. Compartmental organization of projections from the striatum to the substantia nigra in the rat. J Comp Neurol. 1985;236:454–476. 15. Bolam JP, Izzo PN, Graybiel AM. Cellular substrate of the histochemically defined striosome/matrix system of the caudate nucleus: a combined Golgi and immunocytochemical study in cat and ferret. Neuroscience. 1988;24:853–875. 16. Gerfen CR, Staines WA, Arbuthnott GW, Fibiger HC. Crossed connections of the substantia nigra in the rat. J Comp Neurol. 1982;207:283–303. 17. Berendse HW, Voorn P, te Kortschot A, Groenewegen HJ. Nuclear origin of thalamic afferents of the ventral striatum determines their relation to patch/matrix configurations in enkephalinimmunoreactivity in the rat. J Chem Neuroanat. 1988;1:3–10. 18. Gerfen CR. The neostriatal mosaic: compartmentalization of corticostriatal input and striatonigral output systems. Nature. 1984;311:461–464. 19. Donoghue JP, Herkenham M. Neostriatal projections from individual cortical fields conform to histochemically distinct striatal compartments in the rat. Brain Res. 1986;365:397–403. 20. Gerfen CR. The neostriatal mosaic: striatal patch-matrix organization is related to cortical lamination. Science. 1989;246:385–388. 21. Gerfen CR, Engber TM, Mahan LC, Susel Z, Chase TN, Monsma FJ Jr, Sibley DR. D1 and D2 dopamine receptor-regulated gene expression of striatonigral and striatopallidal neurons. Science. 1990;250:1429–1432. 22. Gerfen CR, Young WS 3rd. Distribution of striatonigral and striatopallidal peptidergic neurons in both patch and matrix compartments: an in situ hybridization histochemistry and fluorescent retrograde tracing study. Brain Res. 1988;460:161–167. 23. Pan HS, Penney JB, Young AB. Gamma-aminobutyric acid and benzodiazepine receptor changes induced by unilateral 6-hydroxydopamine lesions of the medial forebrain bundle. J Neurochem. 1985;45:1396–1404.

24. Young WS 3rd, Bonner TI, Brann MR. Mesencephalic dopamine neurons regulate the expression of neuropeptide mRNAs in the rat forebrain. Proc Natl Acad Sci USA. 1986;83:9827–9831. 25. Albin RL, Young AB, Penney JB. The functional anatomy of basal ganglia disorders. Trends Neurosci. 1989;12:366–375. 26. DeLong MR. Primate models of movement disorders of basal ganglia origin. Trends Neurosci. 1990;13:281–285. 27. LeMoine C, Normand E, Guitteny AF, Fouque B, Teoule R, Bloch B. Dopamine receptor gene expression by enkephalin neurons in rat forebrain. Proc Natl Acad Sci USA. 1990;87:230–234. 28. LeMoine C, Bloch B. D1 and D2 dopamine receptor gene expression in the rat striatum: sensitive cRNA probes demonstrate prominent segregation of D1 and D2 mRNAs in distinct neuronal populations of the dorsal and ventral striatum. J Comp Neurol. 1995;355:418–426. 29. Hersch SM, Ciliax BJ, Gutekunst CA, Rees HD, Heilman CJ, Yung KK, Bolam JP, Ince E, Yi H, Levey AI. Electron microscope analysis of D1 and D2 dopamine receptor proteins in the dorsal striatum and their synaptic relationships with motor corticostriatal afferents. J Neurosci. 1995;15:5222–5237. 30. Gong S, Zheng C, Doughty ML, Losos K, Didkovsky N, Schambra UB, Nowak NJ, Joyner A, Leblanc G, Hatten ME, Heintz N. A gene expression atlas of the central nervous system based on bacterial artificial chromosomes. Nature. 2003;425:917–925. 31. Gong S, Doughty M, Harbaugh CR, Cummins A, Hatten ME, Heintz N, Gerfen CR. Targeting Cre recombinase to specific neuron populations with bacterial artificial chromosome constructs. J Neurosci. 2007;27:9817–9823. 32. Bakay RA, DeLong MR, Vitek JL. Posteroventral pallidotomy for Parkinson’s disease. J Neurosurg. 1992;77:487–488. 33. Lozano AM, Lang AE, Hutchison WD, Dostrovsky JO. New developments in understanding the etiology of Parkinson’s disease and in its treatment. Curr Opin Neurobiol. 1998;8:783–790. 34. Birkmayer W, Hornykiewicz O. The L-dihydroxyphenylalanine L-DOPA) effect in Parkinson’s syndrome in man: on the pathogenesis and treatment of Parkinson akinesis. Arch Psychiatr Nervenkr Z Gesamte Neurol Psychiatr. 1962;203:560–574. 35. Bergmann KJ, Mendoza MR, Yahr MD. Parkinson’s disease and long-term levodopa therapy. Adv Neurol. 1987;45:463–467. 36. Gerfen CR, Miyachi S, Paletzi R, Brown P. D1 dopamine receptor supersensitivity in the dopamine-depleted striatum results from a switch in the regulation of ERK1/2MAP kinase. J Neurosci. 2002;22:5042–5054. 37. Berke JD, Paletzki RF, Aronson GJ, Hyman SE, Gerfen CR. A complex program of striatal gene expression induced by dopaminergic stimulation. J Neurosci. 1998;18:5301–5310. 38. Ungerstedt U. Postsynaptic supersensitivity after 6-hydroxydopamine induced degeneration of the nigro-striatal dopamine system. Acta Physiol Scand Suppl. 1971;367:69–93. 39. Robertson GS, Herrera DG, Dragunow M, Robertson HA. L-dopa activates c-fos in the striatum ipsilateral to a 6-hydroxydopamine lesion of the substantia nigra. Eur J Pharmacol. 1989;159:99–100. 40. Robertson GS, Vincent SR, Fibiger HC. Striatonigral projection neurons contain D1 dopamine receptor-activated c-fos. Brain Res. 1990;523:288–290. 41. Graybiel AM, Moratalla R, Robertson HA. Amphetamine and cocaine induce drug-specific activation of the c-fos gene in striosome-matrix compartments and limbic subdivisions of the striatum. Proc Natl Acad Sci USA. 1990;87:6912–6916. 42. Konradi C, Leveque JC, Hyman SEJ. Amphetamine and dopamine-induced immediate early gene expression in striatal neurons

2.1: FUNCTIONAL NEUROANATOMY OF DOPAMINE IN THE STRIATUM

43.

44.

45.

46.

depends on postsynaptic NMDA receptors and calcium. Neuroscience. 1996;16:4231–4239. Keefe KA, Gerfen CR. D1 dopamine receptor-mediated induction of zif268 and c-fos in the dopamine-depleted striatum: differential regulation and independence from NMDA receptors. J Comp Neurol. 1996;367:165–176. Steiner H, Gerfen CR. Cocaine-induced c-fos messenger RNA is inversely related to dynorphin expression in striatum. J Neurosci. 1993;13:5066–5081. Steiner H, Gerfen CR. Dynorphin regulates D1 dopamine receptor-mediated responses in the striatum: relative contributions of pre- and postsynaptic mechanisms in dorsal and ventral striatum demonstrated by altered immediate-early gene induction. J Comp Neurol. 1996;376:530–541. Valjent E, Pascoli V, Svenningsson P, Paul S, Enslen H, Corvol JC, Stipanovich A, Caboche J, Lombroso PJ, Nairn AC, Greengard P, Herve´ D, Girault JA. Regulation of a protein phosphatase cascade allows convergent dopamine and glutamate signals to

47.

48.

49.

50.

21

activate ERK in the striatum. Proc Natl Acad Sci USA. 2005;102:491–496. Gerfen CR, Paletzki R, Worley P. Differences between dorsal and ventral striatum in drd1a dopamine receptor coupling of dopamineand cAMP-regulated phosphoprotein-32 to activation of extracellular signal-regulated kinase. J Neurosci. 2008; 28:7113–7120. Aubert I, Guigoni C, Ha˚kansson K, Li Q, Dovero S, Barthe N, Bioulac BH, Gross CE, Fisone G, Bloch B, Bezard E. Increased D1 dopamine receptor signaling in levodopa-induced dyskinesia. Ann Neurol. 2005;57:17–26. Santini E, Valjent E, Usiello A, Carta M, Borgkvist A, Girault JA, Herve´ D, Greengard P, Fisone G. Critical involvement of cAMP/ DARPP-32 and extracellular signal-regulated protein kinase signaling in L-DOPA-induced dyskinesia. J Neurosci. 2007; 7:6995–7005. Nadjar A, Gerfen CR, Bezard E. Priming for l-dopa-induced dyskinesia in Parkinson’s disease: a feature inherent to the treatment or the disease? Prog Neurobiol. 2009;87:1–9.

2.2

Functional Implications of Dopamine D2 Receptor Localization in Relation to Glutamate Neurons SUS AN R . S ES ACK

INTRODUCTION Dopamine (DA) cell clusters project to large parts of the neural axis, where they provide a critical modulation of diverse functions. Midbrain DA neurons and their projections to forebrain targets have been the subject of extensive research. These cells reside mainly in the substantia nigra pars compacta (SNc) and ventral tegmental area (VTA) and project to widespread areas of the cortex, basal ganglia, limbic forebrain, and thalamus. They sustain tonic rates of firing due to pacemaker potentials and show additional bursts and pauses of activity that reflect their synaptic inputs and signal future expectancy and shifts in attentional resources.1,2 The release of DA in target regions modulates cell excitability in a manner that contributes to motor control, goal-directed behavior, and cognitive function. Malfunctions in midbrain DA neurons and their activity patterns have been implicated in several disease states, including Parkinson’s disease, attention deficit hyperactivity disorder, substance abuse, and schizophrenia.1,3 Among the critical functions of the midbrain DA system is the modulation of glutamate (Glut) transmission at multiple sites where these two transmitter systems interact4,5 (see also Chapter 2.4 by Bolam and Moss and Chapter 7.3 by Surmeier et al. in this volume). As illustrated in Figure 2.2.1, the anatomical connectivity between DA and Glut neurons occurs at three major levels: the dendrites of DA neurons are innervated by Glut axons originating from cortical and subcortical regions (level A); DA axons target the dendrites of Glut neurons primarily in the cortex, hippocampus, amygdala, and thalamus (level B); DA and Glut axons converge onto common target neurons, where they synapse in close proximity (level C). The last configuration occurs at both cortical and basal ganglia sites 22

and facilitates additional presynaptic interactions between DA and Glut. Other, less direct interactions between DA and Glut involve intermediary cell types. In regulating Glut transmission, DA utilizes multiple receptors in two major classes: D1, consisting of D1 and D5 subtypes, and D2, which includes D2, D3, and D4 subtypes. These receptors are extensively expressed at extrasynaptic portions of the plasma membrane, consistent with evidence that DA communicates via volume transmission, in addition to standard synaptic communication.6–8 Consequently, a complete picture of sites where DA and Glut interact can only be achieved by consideration of DA receptor distribution. This review will focus on DA receptors of the D2 class and their spatial and functional relationships with Glut neurons within the circuitry that comprises midbrain DA neurons and their ascending projections to forebrain targets, especially the cerebral cortex and basal ganglia. Interest in D2 receptors has been fueled primarily by their correlation to antipsychotic drug efficacy and their role as autoreceptors. Given the overall similar pharmacology and functions of the D2 receptor class,9 a brief consideration of D3 and D4 receptor subtypes will also be provided. DA also modulates Glut transmission via D1 and D5 receptors, but this subject is beyond the scope of the chapter. The reader is referred to several important papers on the subcellular localization of D1 and D5 receptors in relation to Glut neurons and synapses.10–19 This review will also focus mainly on animal and postmortem human studies. For a review of DA receptor localization in the living human brain and its relationship to Glut function, please see work by Abi-Dargham and Laruelle.20 For in-depth considerations of DA cell anatomy and projections, the reader is referred to comprehensive reviews,21–24 including chapters in this

2.2: FUNCTIONAL IMPLICATIONS OF DOPAMINE D2 RECEPTOR LOCALIZATION

23

FIGURE 2.2.1. Dopamine (DA) acting on D2 receptors (D2Rs) modulates glutamate (Glut) transmission at multiple levels. Solid lines indicate binding of transmitters to their respective receptors; dashed lines indicate secondary actions subsequent to transmitter binding. (A1) Midbrain DA neurons express D2 autoreceptors and receive synaptic Glut afferents. Glut-mediated depolarization would release DA from dendrites to act on D2Rs, which would counteract further excitatory influence from Glut. D2 receptors might also directly modulate Glut transmission at membrane sites where both D2Rs and Glut receptors (GluRs) are distributed in close proximity. (A10 ) A subpopulation of DA neurons coexpresses Glut; these neurons may also contain D2Rs that could modulate Glut release from these neurons. (B2) Dopamine axons directly innervate Glut cells in several target areas, most notably the cortex, amygdala, hippocampus, and thalamus. Modulatory actions on D2Rs at these sites can directly alter the excitability of Glut neurons. (B3, C3) D2 receptors can also influence Glut transmission indirectly via actions on local circuit neurons (e.g., GABA cells in the cortex and cholinergic cells in the striatum). (C4) Dopamine and Glut axons often form closely convergent synapses on target neurons throughout the forebrain, especially the cortex and basal ganglia. Postsynaptic cells expressing D2Rs bind DA following synaptic release or via volume transmission. Activation of D2Rs subsequently modulates Glut transmission, primarily via regulatory actions on AMPA receptors. (A5, C5) Many Glut axon terminals express D2Rs, allowing DA to modulate Glut release presynaptically. (See Color Plate 2.2.1.)

volume by Gerfen (2.1), Haber (2.3), and Moss and Bolam (2.4). More complete consideration of DA receptor subtypes, distribution, and signaling is provided in the chapters by Rankin et al. (3.1) and Fisone (3.3). Many other chapters in this volume cover the physiology and functional importance of midbrain DA systems.

various methods used for these studies. Indeed, the technological state of the art is typically the limiting factor in determining exact receptor distribution, particularly at the cellular and subcellular levels. A full consideration of methodological limitations is beyond the scope of this review and will only be discussed briefly. The reader is encouraged to consult more in-depth reviews of methodology for receptor autoradiography,25,26 in situ hybridization,27,28 and immunocytochemistry.25,29,30

METHODOLOGICAL APPROACHES FOR RECEPTOR LOCALIZATION

Receptor Autoradiography

In considering the anatomical evidence for relationships between D2 receptors and Glut neurons, it is important to keep in mind the advantages and limitations of the

Receptor autoradiography localizes receptors by the binding of radioactive ligands and has the advantage of revealing receptor distribution in brain regions at

24

NEUROANATOMY

high sensitivity regardless of whether or not the receptor is synthesized there. It also can be quantified for determination of changes in receptor density with disease and/or environmental manipulations. However, its usefulness depends on the availability of selective ligands with high affinity and high specific activity, and such ligands are not always available. For example, DA D1 and D5 receptor subtypes have a pharmacology that is too similar to allow their distinction using this approach. Many ligands also bind multiple receptor types, requiring blocking and subtraction methods to distinguish them. Finally, receptor autoradiography allows regional but not cellular or subcellular localization. Selective lesions may help to reveal whether receptors are pre- versus postsynaptic, but interpretation of such experiments is often complicated by compensatory alterations in receptor expression. In Situ Hybridization In situ hybridization identifies the presence of mRNA for a given protein within cells and is therefore essential for determining the capacity of any cell to express a particular receptor. Hybridization probes must be specific for the mRNA species of interest and applied under conditions that prevent nonspecific labeling. Antisense strands tagged with radioisotopes provide a high-sensitivity signal that can be quantified for documenting changes in mRNA levels under natural or experimental conditions. However, radioactive decay reduces the stability of the probes and may necessitate long exposure times for detecting some mRNA species. Digoxigenin or fluorophore tags allow for dual in situ hybridization approaches as well as for combinations of in situ hybridization with immunocytochemistry and with tract tracing. However, these can be less useful for quantitative estimates of mRNA levels. Studies using real-time reverse transcriptase polymerase chain reaction (RT-PCR) indicate that standard in situ hybridization methods are sometimes unable to reveal low levels of receptor mRNA.31 Nevertheless, the results of single-cell RT-PCR experiments also need to be interpreted with caution because amplification may overestimate the actual copy number. In situ hybridization mainly labels cell soma and proximal dendrites where mRNA is most abundant and provides little information regarding the ultimate compartmental destination of receptor proteins following synthesis. Still, in situ hybridization has been used to localize mRNA in distal neuronal compartments,32 although such methodology has not yet been applied to the study of DA receptors.

Immunocytochemistry Immunocytochemistry utilizes antibodies directed against various epitopes of receptor proteins and can detect receptors wherever they are expressed within regions and within cells. The validity of the results is highly dependent on the specificity of antibodies for discrete antigens that are unique to the receptor of interest and are not cross-reacting with other proteins.33 Even when specific reagents are available, different antibodies can suggest different distributions for the same receptor. For reasons that are unclear, some antibodies may preferentially label pre- versus postsynaptic receptors or glial versus neuronal receptors. Hence, combined results using multiple specific antibodies may sometimes be necessary for a complete understanding of receptor distribution. Immunoperoxidase and immunofluorescence methods are ideal for light microscopic studies and are employed with high sensitivity for delineating receptor distribution within cells and neuropil. Efforts to quantify changes in receptor levels following various treatments can be hampered by the amplification methods used to achieve high sensitivity, as well as by variation in tissue fixation, immunoreagent batches, and antibody penetration. Nevertheless, computerassisted image analysis can improve quantification in many cases.25,34 Immunocytochemistry also provides the best approach for subcellular localization of receptor proteins when their exact distribution within neuronal compartments is investigated. It should be noted, however, that when receptors are predominantly trafficked into distal processes, their levels within cell soma can sometimes fall below the detection limits of light microscopy. This has often been the case for DA receptors, making it difficult to apply dual immunolabeling or tract tracing to the phenotypic identification of cells that express these receptors. The use of confocal microscopy can improve detection of low DA receptor levels in soma,35,36 and electron microscopy also permits detection of sparse immunoreactivity in perikarya as well as distal processes.15,37,38 Two additional methodological issues affect the interpretation of subcellular localization studies of DA receptors. First, there is often a trade-off between sensitivity and spatial resolution. Specifically, immunoperoxidase methods permit only a relative illustration of subcellular distribution because the reaction product is diffusible. Nevertheless, these approaches typically employ signal amplification, giving them the sensitivity to detect receptors present at low levels. Immunogold techniques use a nondiffusible marker and so have better spatial resolution for more precise subcellular

2.2: FUNCTIONAL IMPLICATIONS OF DOPAMINE D2 RECEPTOR LOCALIZATION

localization, although this is often achieved at the expense of sensitivity.30 Second, pre-embedding immunogold methods tend to bias detection toward extrasynaptic receptors, in part because of limited antibody penetration of the protein complex at synaptic junctions. Postembedding techniques involve application of antibodies after ultrathin sectioning and therefore afford better penetration of synapses. Nevertheless, the processes required for plastic embedding (especially lipid fixation) often have the effect of destroying antigenicity. Such limitations can sometimes be overcome, as they were for an elegant series of studies on Glut receptor localization by Somogyi and colleagues.39 Unfortunately, the precise distribution of DA receptors with respect to synapses has yet to be accomplished with similar methodology. Hence, it is important to keep in mind while reading this review that the exact spatial location of D2 receptors with respect to Glut or other synapses is not yet known, although some speculative observations on this subject are occasionally offered.

D2 RECEPTOR DISTRIBUTION AND FUNCTION Midbrain DA Neurons D2 autoreceptors on DA neurons D2 autoreceptors expressed by midbrain DA neurons are activated by dendritically released DA and moderate cell activity in the face of phasic excitation40 driven by Glut afferents from multiple sources.41 In addition to counteracting excitatory drive, D2 receptors occur in close proximity to the synapses formed by Glut axons and can therefore modulate transmission via actions on Glut receptors (Fig. 2.2.1, site 1). Numerous anatomical studies document the expression of D2 receptors in the ventral midbrain.9,13,37,38,42–46 D2 receptor protein has been specifically localized to DA cells,13,38 with occasional non-DA neurons also appearing to express this receptor.38,47 One study suggests that D2 receptors are not uniformly expressed by all populations of DA cells, being notably lower in the parabrachial than the paranigral division of the VTA.47 Although retrograde tract tracing was not employed in this study, projections to the prefrontal cortex (PFC) originate more commonly from the parabrachial subdivision, making these observations consistent with physiological and neurochemical reports that mesocortical DA neurons are less sensitive to autoreceptor control than mesolimbic or nigrostriatal cells.48 For many years, midbrain DA neurons were thought to receive Glut afferents from a limited number of sources, most prominently the PFC and brainstem

25

mesopontine tegmentum. More recently, the seminal work of Geisler and colleagues has revealed numerous sources of Glut afferents to the VTA that, in addition to the cortex, arise from widespread subcortical structures including the lateral hypothalamic and lateral preoptic areas, central gray, raphe nuclei, brainstem tegmentum, and reticular formation.41 Additional Glut afferents to the SNc originate from the subthalamic nucleus.49 Our own ultrastructural studies have verified that the majority of Glut axons in the VTA originate from probable subcortical as opposed to cortical sites.50 Collectively, DA neurons receive relatively abundant synaptic input from these various Glut afferents, although SNc DA cells exhibit synaptic input more commonly from GABA axons.51 Dopamine cells also express ionotropic and metabotropic Glut receptors, in some cases near sites of presumed Glut synapses.52–54 Within midbrain DA and non-DA cells, the majority of D2 receptor immunoreactivity has been localized to the cytoplasm along the surface of smooth endoplasmic reticulum as well as to the plasma membrane at extrasynaptic sites.13,38,47 The D2 receptor has occasionally been detected near symmetric or asymmetric synapses when the subcellular distribution was examined using immunoperoxidase methods. Immunogold localization has verified the cytoplasmic and endosomal labeling as well as the distribution of D2 receptors to extrasynaptic portions of the plasmalemma. However, punctate gold labeling near asymmetric synapses has not yet been observed,13 suggesting that autoreceptors may not directly modulate Glut transmission in the SNc or VTA. Moreover, there are as yet no studies examining the proximity of D2 receptors to N-methyl-D-aspartate (NMDA), AMPA, or metabotropic Glut receptors in DA neurons. Nevertheless, some electrophysiological studies do report modulatory interactions. Specifically, long-term depression (LTD) in VTA DA neurons can be completely blocked by amphetamine via a D2-receptordependent mechanism.55,56 Such an effect may consequently result in the expression of long-term potentiation (LTP) and excessive Glut transmission with chronic amphetamine abuse.55 Moreover, it has been conjectured that blocking D2 receptors with antipsychotic drugs may contribute to the treatment of schizophrenia by facilitating the development of LTD.55 The exact mechanism whereby D2 receptors alter Glut plasticity in DA cells remains to be determined. Possible D2 autoreceptors on DA neurons that colocalize Glut Dopamine and Glut also reportedly intersect within DA neurons in that some DA cells have been reported to

26

NEUROANATOMY

colocalize Glut. An in-depth evaluation of this subject is beyond the scope of this chapter; nevertheless, a few comments are in order. Early reports of extensive colocalization of Glut in most DA neurons were based on indirect measures or on analysis of cultured cells.57–59 In addition, many indications of a Glut phenotype have been complicated by the presence of this substance for metabolic purposes or as a precursor for GABA. More recently, several laboratories have reported a substantial population of VTA (but not SNc) neurons expressing mRNA for the vesicular glutamate transporter type 2 (VGlut2), an accepted marker of Glut cells.7,60–62 Most midbrain Glut neurons appear to be a separate population from DA cells, although there is some colocalization. Standard in situ hybridization estimates in adult animals range from 2% or less7,61,62 to as high as 20%–50% in certain midline divisions60 (but see 61,62). Single-cell RTPCR experiments suggest that 25% of DA neurons may express a low abundance of VGlut2 mRNA.7 This coexpression is developmentally regulated and largely suppressed in adulthood, although it might be reinduced under pathological conditions.7 In any case, Glut appears to be colocalized only in a subset of DA neurons. Regarding the targets of VTA cells containing both DA and Glut, at least some of these neurons are reported to project to the core of the NAc7 but not to the dorsal striatum.63 VGlut2-containing VTA cells also project to the PFC,64 but it has not yet been determined whether any of these neurons coexpress DA. A preliminary report indicates that Glut VTA cells have local collaterals that innervate DA neurons.65 Hence, Glut afferents to DA cells derive from intrinsic as well as extrinsic sources. It is unclear how the combined expression of DA and Glut interacts functionally within target regions, although some theoretical models have been generated.7,66,67 In future studies, it will be important to determine the extent to which D2 autoreceptors at soma (Fig. 2.2.1, site 10 ) or nerve terminals are involved in regulating Glut corelease. Such an action has been suggested from cultured cell studies.58 Forebrain Glut Cells D2 receptors on Glut neurons Dopamine cells directly innervate Glut neurons within the cerebral cortex, hippocampus, amygdala, and thalamus68–71 (Fig. 2.2.1, site 2). The cortical DA projection is densest in frontal areas, including various divisions of the medial PFC, orbital cortex, and cingulate cortex. A projection to the motor cortex is prominent in primates but only modest in the rat.23 Dopamine axons also innervate Glut cells in the subthalamic

nucleus,72,73 but this structure expresses mainly D5 receptors74 and will not be considered further here. D2 receptors as measured by receptor autoradiography or in situ hybridization are expressed in the cerebral cortex at levels noticeably lower than in the basal ganglia and are also lower than the dominant D1 subtype.9,42,45,75–78 D2 receptor binding and/or mRNA have also been described in the amygdala, hippocampus, and thalamus.9,79,80 In multiple cortical regions, and especially in the PFC and cingulate cortex, D2 mRNA is expressed most heavily in the pyramidal cell layers 2–3 and 5–6,75–78 and specifically in neurons projecting to the striatum and to other cortical regions.75 The results of immunohistochemical localization studies are relatively well matched to the distribution of D2 receptor mRNA.15,37,43,46,80–83 In addition, antibodies have identified D2 receptors within astrocytic processes, where they reportedly increase calcium levels.82,84 Whether this ultimately provides a mechanism for regulation of Glut transmission85 has not yet been determined. A few ultrastructural studies using immunogold methods have described the subcellular distribution of D2 receptor immunoreactivity in the cortex, which is found most commonly in the cytoplasm associated with endosomes, smooth endoplasmic reticulum, or Golgi apparatus, and along the plasma membrane at nonsynaptic sites.15,82,84,86–88 D2 receptors are also frequently associated with clathrin-coated vesicles, suggesting that these organelles form an important component of recycling for this receptor.87 To date, none of these studies has localized D2 receptors by immunogold in close proximity to asymmetric, presumed Glut synapses or to identified Glut receptors on the dendrites of cortical cells. Nevertheless, considerable functional interactions between D2 and Glut receptors have been documented. Many of these interactions exemplify DA regulation of Glut transmission via convergence onto common targets (see the section ‘‘Functional Implications of DA and Glut Convergence’’ below). However, as the target neurons are themselves glutamatergic, the interactions will be discussed here. Electrophysiological evidence, particularly from the PFC, suggests that DA acting on D2 receptors decreases the excitability of cortical pyramidal cells and suppresses Glut synaptic responses, especially those of AMPA receptors89–92 (see also 93). In other cortical regions, the direction of D2 modulation of Glut transmission varies from attenuating to facilitating.81,94–96 These variable responses appear to depend on multiple factors, including target region, cortical layer, the activity state of neurons, and timing issues relative to short-term plasticity.5,97

2.2: FUNCTIONAL IMPLICATIONS OF DOPAMINE D2 RECEPTOR LOCALIZATION

Within the PFC, the actions of D2 receptors on Glut cells are often opposite those of D1 receptors, which typically increase cell excitability and Glut transmission.5,90,91,93,98,99 D2 receptors also alter plasticity at pyramidal neurons by favoring LTD of Glut transmission100; D1 receptors again have opposing actions by facilitating LTP.93,99,101 Collectively, the marked differences in D2 versus D1 receptor actions in the PFC have important implications for the working memory functions of this region.5 D2 receptors on secondary neurons The activity patterns of cortical Glut pyramidal cells are strongly influenced by GABA local circuit neurons,102 which are also innervated by DA axons103,104 and express D2 receptors.46,84 Hence, non-Glut neurons constitute another site at which DA can influence Glut transmission, albeit indirectly (Fig. 2.2.1, site B3). As described above, the dominant action of D2 receptors on cortical Glut neurons is suppression of excitability and synaptic responses. This suppressive effect is likely to be accomplished, at least in part, by increasing GABAergic drive onto these cells.92,105–107 Interestingly, the strength and duration of this effect increase in postpubertal animals.92,107 Secondary neurons in other target areas also mediate some of the indirect actions of D2 receptors on Glut transmission,5 although many of these are mediated at the presynaptic level (see below). Convergence of DA and Glut D2 receptors on the postsynaptic targets of DA and Glut inputs Throughout their target fields, DA axons synapse onto neurons that also receive convergent Glut afferents, providing one of the main sites where these two neuroregulators interact (Fig. 2.2.1, site 4). Within the cortex, these convergence sites include mainly the spines of Glut pyramidal neurons,68 although DA synapses onto local circuit neurons also occur in close proximity to Glut contacts.103 Within the basal ganglia, convergence sites for DA and Glut axons are primarily the spines and distal dendrites of GABA projection neurons.72,108,109 For a more thorough description of the anatomy of DA and Glut convergence, the reader should consult Chapter 2.4 by Moss and Bolam in this volume. Functional sites where DA and Glut can interact via postsynaptic receptors range from highly compartmentalized in the case of spine convergence110 to more diffuse in instances where DA receptors alter overall

27

cell excitability regardless of the focal sites where Glut is released and initiates depolarization.111,112 Evidence for D2 receptor expression in cortical structures was previously considered in the section ‘‘Forebrain Glut Cells,’’ so this section will focus on D2 receptor distribution within basal ganglia regions. Cellular localization studies. The presence of DA D2 receptors throughout the dorsal to ventral aspects of the striatopallidal complex has been amply demonstrated by receptor autoradiography, in situ hybridization, and immunohistochemical methods.8,9,13,37,42–46,76,113–115 D2 receptors are expressed by cholinergic and neurotensin interneurons116–118 and GABAergic medium spiny neurons.13,114,117,119–121 The D2 receptor has also occasionally been identified within glial processes in the striatum and pallidum.38,115 Within the striatum, the D2 receptor is localized mainly to cells that express enkephalin and project to the external globus pallidus (i.e., the indirect striatal output pathway).36,117,119,120,122,123 Most of these neurons do not express detectable levels of D1 receptor and therefore do not appear to contribute substantially to the direct output pathway projecting to the substantia nigra.12,120,124 Nevertheless, recent studies indicate a greater degree of collateralization in both direct and indirect striatal output channels than was previously appreciated.125 Moreover, mRNA amplification indicates that nearly half of striatal medium spiny neurons have the capacity to express a combination of DA receptor subtypes when one includes the extended D1– D5 and D2–D3–D4 families.31 This finding does not necessarily mean that protein for multiple receptor types is expressed at functional levels. Indeed, early quantification revealed little evidence for receptor coexpression within striatal spines.12 Nevertheless, confocal microscopic studies suggest a greater extent of colocalization within soma, and physiological evidence indicates that many striatal neurons show comparable responses to D1- and D2-selective agonists.31,35,122,123 Such conflicting reports have raised questions regarding the sensitivity and accuracy of various methodologies for estimating receptor expression. Some resolution of this controversy appears to come from BAC transgenic mouse strains in which expression of enhanced green fluorescent protein (EGFP) is selectively linked to the promoter for either D1 or D2 receptors. In these animals, the D1 and D2 subtypes are largely segregated in separate medium spiny cell populations.126–128 Moreover, experiments in BAC transgenic mice have begun to reveal some of the complex mechanisms whereby D1- or D2-selective agonists can induce comparable physiological actions in direct

28

NEUROANATOMY

and indirect pathway cells. For example, D2 receptors on cholinergic interneurons reduce acetylcholine release, which triggers a cascade of events that ultimately attenuates Glut release onto principal cells, regardless of whether they express D1 or D2 receptors128 (Fig. 2.2.1, site C3). Such a mechanism serves as a reminder that indirect actions of D2 receptors can occur even when the Glut neuronal elements do not express these receptors or lie in close proximity to the site of DA release. Subcellular localization studies. Within spiny striatal neurons, some studies have noted that immunoreactivity for the D2 receptor is more intense in distal portions of the dendritic tree than in soma and proximal dendrites.12,37,38,129 This distribution is consistent with physiological studies emphasizing the dominant functionality of distal DA receptors96,130,131 and with the location of Glut synapses and receptors in dendrites.132,133 The subcellular distribution of the D2 receptor is similar to that reported in the cortex (see above), namely, being cytoplasmic in association with organelles (e.g., endosomes and Golgi apparatus) and plasmalemmal at nonsynaptic locations.13,86,88,115,134 The proportion of membrane to cytoplasmic receptor appears to increase from the soma outward into the dendritic tree.134 In addition to extrasynaptic sites, D2 receptor immunolabeling occurs near asymmetric and symmetric synapses,12,14,37,38,135 and this distribution has been confirmed by discrete immunogold methods.13 D2 receptors near asymmetric synapses provide potential anatomical substrates for modulation of Glut transmission at these sites. This interpretation is supported by observations of D2 receptor immunolabeling within spines that receive synapses from axons labeled for Glut14 or tracer transported anterogradely from the motor cortex12 or the PFC.135 Receptors associated with symmetric synapses might be postsynaptic to DA axons, although this has not yet been demonstrated directly. Some supportive evidence for D2 postsynaptic receptors does exist (Sesack, unpublished observations), and functional studies have speculated that these are the logical subtype to receive synaptic DA signals.136 D2 receptors also occur within dendritic structures receiving symmetric synaptic input from GABA-labeled terminals.14,121 Functional implications of DA and Glut convergence. Sites where DA and Glut axons synapse onto the same spines place DA in a position to modulate postsynaptic responses to highly specific sources of Glut drive. This ‘‘triadic’’ configuration also places DA

in a favorable position for reaching extrasynaptic heteroreceptors on Glut nerve terminals (see below). Spines are now understood to be biochemically rather than electrically isolated compartments. Diffusion of large signaling molecules is relatively restricted, allowing individual spines with specific synaptic inputs to undergo selective alterations in activity and plasticity.110 Moreover, evidence suggests that the movement of membrane-bound receptors (at least for Glut) is tightly regulated within and between spines in an activity-dependent manner.137 How DA receptors are trafficked in accordance with such phenomena or themselves contribute to the movement of Glut receptors is only beginning to be explored.93,110,138 The interesting speculation has been put forward that spines with Glut synapses but no converging DA synapse (i.e., dyads versus triads) may be the main detectors of increased DA levels, assuming that they express DA receptors.110 In this case, DA modulation of Glut transmission would occur only when extracellular levels of DA rise, either from increased tonic activity or short-term phasic bursts. Moreover, depending on the distance from the focal source of DA release, each spine could be set to different levels of activity and plasticity. Hence, spines that express D2 receptors do not necessarily need to receive DA from a synaptic release event. In this regard, it is interesting to note evidence suggesting that DA axons form a regular lattice in the striatum that maintains relatively constant spacing (~1 mm) from thalamostriatal and corticostriatal Glut axons112 (see also Chapter 2.4 by Moss and Bolam in this volume). Moreover, it has been estimated that DA release from synaptic or extrasynaptic sites can diffuse 7–8 mm and still be in sufficient concentration to stimulate D2 receptors.8 Acute D2 modulation of Glut transmission is believed to occur primarily through regulation of AMPA receptors, which are located in distal dendrites and spines in striatal neurons.133 Specifically, D2-induced reduction of membrane AMPA receptors decreases the excitability of neurons.4,130,131 D2 receptors have additional actions that further reduce excitability and the response to synaptic Glut, including indirect effects mediated by reducing acetylcholine release from cholinergic interneurons.131 As was found in the cortex, the actions of striatal D2 receptors are typically opposite the actions of D1 receptors that facilitate Glut transmission.4,131 Interestingly, D1 and D2 receptor–expressing cell populations also appear to receive Glut input from different cortical sources.139 It is likely that D2-mediated reduction of the corticostriatal Glut drive serves to suppress the selection of inappropriate motor programs elicited by uncoordinated cortical events.131 However, striatal

2.2: FUNCTIONAL IMPLICATIONS OF DOPAMINE D2 RECEPTOR LOCALIZATION

neurons expressing D2 receptors have recently been shown to be more excitable than those expressing D1 receptors and to more faithfully represent the corticostriatal drive.140 These characteristics appear to be independent of D2 and AMPA receptors, and how they contribute to adaptive motor control remains to be established. Structural plasticity regulated by D2 receptors. Longterm structural plasticity at the spines of striatal medium spiny neurons represents an important functional event that reflects DA interactions with Glut via D2 receptors. Many studies document the importance of DA receptors for synaptic plasticity, with D2 receptors favoring LTD and D1 receptors facilitating LTP131,138,141 (see also Chapter 7.3 by Surmeier et al. in this volume). That DA is at least partly responsible for structural maintenance in the striatum is evidenced by reports of 27% spine loss in postmortem Parkinson’s brain142 and 15%–30% reduction in spines and axo-spinous synapses in animal models with DA denervation.143,144 The degree of spine loss is tightly correlated to the reduction in DA fibers,144 and lesion-induced spine loss is also noted in the nucleus accumbens (NAc) and PFC.145 Disruptions of other aspects of dendritic morphology have also been reported following DA depletion in animal models and in Parkinson’s disease (for a review, see 73). A recent collaborative study indicates that loss of axospinous synapses induced by DA denervation actually reflects a higher-magnitude reduction (~35%–50%) in the population of neurons expressing the D2 receptor combined with no change in D1 receptor–expressing cells.126 Both the spines themselves and the Glut synapses innervating the heads of spines are lost, and postsynaptic excitatory potentials are also substantially reduced. These findings suggest that D2 receptors stabilize both the pre- and postsynaptic elements in corticostriatal or thalamostriatal axo-spinous synapses on striatopallidal neurons. Reduced spine density has also been observed with reserpine treatment that merely depletes DA,126 suggesting that it is the loss of DA itself that destabilizes spines and not the physical removal of the nerve terminal. This is consistent with the well documented ability of DA to communicate via extrasynaptic receptors.6 The exact mechanism linking D2 receptors to spine stabilization has not yet been fully elucidated. D2 receptor activation can produce antioxidant effects and protect against Glut cytotoxicity.146,147 However, the physiological studies of Surmeier indicate that D2 receptor–mediated spine stabilization depends at least in part on selective inhibition of L-type calcium

29

channels with a Cav1.3a1 subunit.126 This channel is expressed in spines and linked to Glut synapses by scaffolding proteins. Loss of D2 inhibitory influence reduces spines and Glut axo-spinous synapses but leads conversely to compensatory increases in Glut-driven activity in striatopallidal cells that may contribute to the symptoms of Parkinson’s disease. This mechanism might be specific for the striatum, as a similar loss of spines in the PFC following DA lesioning appears not to be linked to D2 receptors.145,148 The selective spine stabilization by D2 receptors in the striatum may also be specific to rodents, given that DA denervation in the primate basal ganglia produces spine loss that is not selective for D2-expressing cells.144 However, it should be noted that the latter study did not employ unbiased stereological measurements. Despite reducing the density of excitatory axo-spinous synapses overall, selective DA lesions actually increase the number of complex synapses with perforations,143 a morphological feature thought to reflect enhanced synaptic efficacy.149,150 This effect clearly involves synapses formed by Glut axons and is mimicked by chronic treatment with D2 receptor antagonists, suggesting a role in the neurological side effects produced by antipsychotic drugs.151 Nevertheless, the exact contribution of D2 receptors to this phenomenon is unclear.152 Changes in spine density in various forebrain regions also accompany behavioral sensitization to chronic psychostimulants that enhance DA levels.153 However, a clear correlation to D2 receptor activation also has not yet been demonstrated.127 D2 heteroreceptors on Glut nerve terminals Glut neurons that express D2 receptors have the potential to transport these receptors into axons, where they can be inserted into the presynaptic membrane (Fig. 2.2.1, sites A5, C5). Although DA and Glut axons do not exhibit axo-axonic synapses, DA can reach presynaptic heteroreceptors via diffusion over distances of several microns.8,112 Moreover, the ability of DA to act on these receptors is likely to be facilitated by the frequent close apposition of DA and Glut axons and their synaptic convergence onto common distal dendrites, as observed in the cortex and striatum (see above). Although presynaptic D1 receptors are sometimes described on Glut nerve terminals,16 it is the D2 class that is the most common presynaptic receptor type.154 Forebrain. Presynaptic D2 receptors have been reported in numerous brain regions, including the striatal complex, ventral pallidum, amygdala, and cerebral Within cortex.12,13,23,37,82,83,87,88,114–116,121,129,135,139

30

NEUROANATOMY

axons, D2 receptor immunolabeling is distributed to both cytoplasmic and plasmalemmal sites.82,87,135 D2 receptor–bearing varicosities forming asymmetric synapses presumably represent Glut nerve terminals. Although the presence of Glut has not been directly demonstrated in these profiles, some have been shown to originate from the PFC.83,135 These observations are consistent with physiological demonstrations of presynaptic inhibition of Glut transmission via D2 receptors on axons from the PFC and other Glut sources.154–158 Other presynaptic D2 receptors represent autoreceptors on DA nerve terminals38,115 or heteroreceptors regulating the release of other transmitters like GABA.121 AMPA receptors clearly serve as presynaptic autoreceptors in striatal structures.63 Hence, the functional interactions reported between D2 and AMPA receptors in the soma and dendrites of cortical and striatal cells4,130,131 (see above) may also play out within axon terminals. However, to date, no study has endeavored to localize both D2 heteroreceptors and AMPA autoreceptors within the same axon terminals. In the striatum, the extent of D2 inhibition of Glut release increases with the firing frequency of cortical afferents and is selective for terminals with low release probability.154 In this way, DA acts as a low-pass filter to favor corticostriatal transmission via the most active synaptic connections. Although the ability of D2 receptors to reduce Glut release is likely to occur through presynaptic receptors, studies also suggest that this effect may involve a postsynaptic D2 receptor action and retrograde signaling via endocannabinoids159 (see also 128). Midbrain. Within the VTA, D2 receptor labeling has also been reported in axons, most of which do not form synapses in single sections.38,47 Some axon varicosities containing weak immunolabeling for D2 receptor and forming asymmetric synapses have been described.47 This observation is consistent with one electrophysiological study reporting presynaptic inhibition of Glut transmission by D2 receptors in the VTA.160 The latter authors have speculated that dendritically released DA might act on these receptors in order to reduce the Glut drive associated with burst firing.

D3 RECEPTORS IN RELATION TO GLUTAMATE The D3 receptor subtype has garnered extensive interest in neuropsychopharmacology research because of its relatively restricted expression to the ventral parts of the striatal complex and to the cortex. This distribution opens up the possibility that novel D3-selective antipsychotics might be free of the adverse effects mediated by

D2 receptors in dorsal striatal regions.161–163 The D3 receptor subtype has also been suggested as a potential therapeutic target for the treatment of drug addiction.164 D3 receptors have been localized by receptor autoradiography, in situ hybridization, and immunocytochemistry. Despite some discrepancies among the three approaches, there is good agreement that the most robust D3 expression is in the ventral striatal complex, including the NAc shell, olfactory tubercle, and islands of Calleja.9,42,46,124,161,163–169 Lower levels of D3 receptors have also been reported in the caudate putamen, ventral pallidum, hippocampus, amygdala, and midbrain. In striatal neurons, the D3 receptor is at least partially colocalized with D1 and D2 receptors and is most abundant in neurons that coexpress substance P and project to the substantia nigra.31,124,165,166 Many investigators report the presence of D3 receptors in the cerebral cortex, most notably the parietal region and PFC.46,163,165,167,168 D3 receptors localized to layer 3– 5 cells may be present in Glut pyramidal neurons, although many may be in nonpyramidal cells.46 A fair amount of controversy surrounds the issue of whether midbrain D3 receptors represent actual autoreceptors. Estimates range from weak expression in small populations of DA cells to nearly complete expression in all DA neurons.166,168 Immunoreactivity also occurs within non-DA cells and in the neuropil of the substantia nigra, which may correspond in part to presynaptic receptors on striatonigral axons.165,168,169 Selective lesions of DA neurons reduce D3 receptor binding in the NAc, but they also reduce mRNA for this receptor, suggesting that an anterograde factor in DA axons may be necessary for sustaining postsynaptic D3 receptors.162,169 Further challenges to the establishment of D3 autoreceptors come from transgenic mouse models, where knockouts of D3 receptors produce little change in autoreceptor functions, whereas D2 knockouts eliminate evidence of autoreceptor activity.40,170 Most recently, it has been suggested that D3 autoreceptors control only basal levels of DA and not release or firing activity.171 Compared to the other D2 family receptors, much less is known about the subcellular localization of the D3 subtype. Ultrastructural studies are especially lacking, perhaps due to concerns regarding antibody specificity.46,165,168 Where it has been estimated from light microscopy, the D3 receptor has been described as occurring in the cytoplasm and at extrasynaptic portions of the plasma membrane.168 Given the limited subcellular information, one can only speculate regarding potential sites where D3 receptors might interact with Glut transmission: (1) as potential

2.2: FUNCTIONAL IMPLICATIONS OF DOPAMINE D2 RECEPTOR LOCALIZATION

autoreceptors on some midbrain DA neurons that receive Glut input (or colocalize Glut), (2) as postsynaptic receptors on cortical pyramidal or non-pyramidal cells, and (3) as postsynaptic receptors on ventral striatal neurons that occur in proximity to Glut inputs. Clearly, there is a need for more detailed studies of D3 receptor localization before more specific relationships to Glut transmission can be established.

D4 RECEPTORS IN RELATION TO GLUTAMATE There is considerable interest in the localization of the DA D4 receptor because of its higher affinity for the atypical antipsychotic drug clozapine and the potential contribution of this receptor subtype to the understanding of schizophrenia pathophysiology and treatment.172,173 D4 receptors have also been linked to attention deficit hyperactivity disorder and the regulation of novelty seeking.174 The DA D4 receptor is expressed predominantly in cortical regions, suggesting a particular role in the modulation of cognitive functions. However, the various approaches for detecting D4 receptors have produced somewhat conflicting results. By in situ hybridization or immunocytochemistry, the density of D4 receptor expression is as follows: densest in motor, sensory, and cingulate cortices; intermediate in temporal, retrosplenial, and granular association areas; and lowest in the PFC and cortical regions near the rhinal sulcus. D4 receptors have also been observed in the hippocampus and amygdala.9,46,76,175–179 D4 receptor mRNA is reportedly concentrated in deep cortical layers, especially layer 5.76–78 Immunoreactivity for the D4 receptor protein is densest in layers 2–5, including a modest expression in Glut pyramidal cells and dense expression in GABA local circuit neurons.46,173,176,179 GABA neurons in the striatum and pallidum also appear to express measurable levels of D4 receptor mRNA or protein, although these have been difficult to detect in some studies.31,46,173,176,178,180 Recently, BAC transgenic mice have been used to selectively express EGFP in cells transcribing the D4 receptor gene.181 The densest expression was reported in deep layers of the PFC, including prelimbic, cingulate, orbital, and agranular insular areas. D4 receptors were also observed in motor and piriform cortices, the anterior olfactory nucleus, ventral pallidum, and brainstem parabrachial nucleus. No signal was detected in the striatum or NAc, amygdala, hippocampus, thalamus, or midbrain. It is not yet known whether the discrepancies in D4 receptor expression between BAC transgenic mice and other anatomical localization methods are due

31

to sensitivity issues, problems with the specificity of some D4 receptor antibodies, or species differences. At the subcellular level in both the cortex and striatum, immunoreactivity for the D4 receptor has been reported along nonsynaptic portions of the plasma membrane throughout soma, dendrites, and spines; receptor localization to the smooth endoplasmic reticulum and other cytoplasmic organelles has also been described.173,180 The dominant nonsynaptic distribution is consistent with observations that D4 receptors occur at considerable distances from DA axons179 (but see 181). D4 receptor immunoreactivity has specifically been localized in proximity to Glut synapses in NAc spines by both immunoperoxidase and immunogold methods,180 suggesting that the D4 receptor may directly modulate Glut transmission at these sites. As yet, there are no physiological studies to support this, although there is evidence for D4 receptor-mediated attenuation of Glut transmission in the amygdala and PFC.182–184 D4 receptors may also reduce transmission through GABA-A receptors in the PFC.185 The anatomical substrates for these interactions have not yet been explored. The majority of D4 receptors in the striatal complex appear to be presynaptic.180 These D4 immunoreactive axons either fail to form synapses in single sections, precluding their identification, or make asymmetric contacts characteristic of Glut connections. The source of these latter axons is not known but is most likely the cerebral cortex.177,180 The thalamus is another potential source, but thalamic D4 receptors are primarily expressed by intrinsic reticular cells173 and not projection neurons. It should be noted that some cross-reaction of the D4 antibody with the D2 receptor could not be ruled out in the NAc ultrastructural study.180 This is important to consider given that much of the distribution reported for D4 receptors has also been observed in D2 localization studies. Importantly, Glut responses seem to be unaltered in the striatum of mice with transgenic deletion of D4 receptors, whereas D2 receptor knockout mice do show substantial changes in excitatory synaptic activity.157 This suggests that D2 receptors may be the dominant presynaptic controller of Glut transmission, at least in the striatum. In the cortex, D4 receptors do have presynaptic actions on Glut axons,186 consistent with the greater density of D4 receptors there compared to the striatum. In summary, the main sites of probable DA modulation of Glut transmission via the D4 receptor are (1) directly onto Glut pyramidal neurons of PFC and other cortical regions, (2) indirect regulation of pyramidal neurons via actions on GABA local circuit neurons, (3) modulation of Glut transmission in NAc spines, and (4) presynaptic actions on Glut axons in the NAc.

32

NEUROANATOMY

REFERENCES 1. Schultz W. Multiple dopamine functions at different time courses. Annu Rev Neurosci. 2007;30:259–288. 2. Redgrave P, Gurney K, Reynolds J. What is reinforced by phasic dopamine signals? Brain Res Rev. 2008;58(2):322–339. 3. Liss B, Roeper J. Individual dopamine midbrain neurons: functional diversity and flexibility in health and disease. Brain Res Rev. 2008;58(2):314–321. 4. Cepeda C, Buchwald NA, Levine MS. Neuromodulatory actions of dopamine in the neostriatum are dependent on the excitatory amino acid receptor subtypes activated. Proc Natl Acad Sci USA. 1993;90(20):9576–9580. 5. Seamans JK, Yang CR. The principal features and mechanisms of dopamine modulation in the prefrontal cortex. Prog Neurobiol. 2004;74(1):1–58. 6. Zoli M, Torri C, Ferrari R, et al. The emergence of the volume transmission concept. Brain Res Rev. 1998;26(2-3): 136–147. 7. Descarries L, Berube-Carriere N, Riad M, Bo GD, Mendez JA, Trudeau LE. Glutamate in dopamine neurons: synaptic versus diffuse transmission. Brain Res Rev. 2008;58(2):290–302. 8. Rice ME, Cragg SJ. Dopamine spillover after quantal release: rethinking dopamine transmission in the nigrostriatal pathway. Brain Res Rev. 2008;58(2):303–313. 9. Missale C, Nash SR, Robinson SW, Jaber M, Caron MG. Dopamine receptors: from structure to function. Physiol Rev. 1998;78(1):189–225. 10. Smiley JF, Levey AI, Ciliax BJ, Goldman-Rakic PS. D1 dopamine receptor immunoreactivity in human and monkey cerebral cortex: predominant and extrasynaptic localization in dendritic spines. Proc Natl Acad Sci USA. 1994;91(12):5720–5724. 11. Bergson C, Mrzljak L, Smiley JF, Pappy M, Levenson R, Goldman-Rakic PS. Regional, cellular, and subcellular variations in the distribution of D1 and D5 receptors in primate brain. J Neurosci. 1995;15(12):7821–7836. 12. Hersch SM, Ciliax BJ, Gutekunst CA, et al. Electron microscopic analysis of D1 and D2 dopamine receptor proteins in the dorsal striatum and their synaptic relationships with motor corticostriatal afferents. J Neurosci. 1995;15(7 pt 2):5222–5237. 13. Yung KK, Bolam JP, Smith AD, Hersch SM, Ciliax BJ, Levey AI. Immunocytochemical localization of D1 and D2 dopamine receptors in the basal ganglia of the rat: light and electron microscopy. Neuroscience. 1995;65(3):709–730. 14. Yung KK, Bolam JP. Localization of dopamine D1 and D2 receptors in the rat neostriatum: synaptic interaction with glutamate- and GABA-containing axonal terminals. Synapse. 2000;38(4):413–420. 15. Paspalas CD, Goldman-Rakic PS. Microdomains for dopamine volume neurotransmission in primate prefrontal cortex. J Neurosci. 2004;24(23):5292–5300. 16. Paspalas CD, Goldman-Rakic PS. Presynaptic D1 dopamine receptors in primate prefrontal cortex: target-specific expression in the glutamatergic synapse. J Neurosci. 2005;25(5):1260–1267. 17. Hara Y, Pickel VM. Overlapping intracellular and differential synaptic distributions of dopamine D1 and glutamate N-methylD-aspartate receptors in rat nucleus accumbens. J Comp Neurol. 2005;492(4):442–455. 18. Pickel VM, Colago EE, Mania I, Molosh AI, Rainnie DG. Dopamine D1 receptors co-distribute with N-methyl-D-aspartic acid type-1 subunits and modulate synaptically-evoked Nmethyl-D-aspartic acid currents in rat basolateral amygdala. Neuroscience. 2006;142(3):671–690.

19. Bordelon-Glausier JR, Khan ZU, Muly EC. Quantification of D1 and D5 dopamine receptor localization in layers I, III, and V of Macaca mulatta prefrontal cortical area 9: coexpression in dendritic spines and axon terminals. J Comp Neurol. 2008; 508(6):893–905. 20. Abi-Dargham A, Laruelle M. Mechanisms of action of second generation antipsychotic drugs in schizophrenia: insights from brain imaging studies. Eur Psychiatry. 2005;20(1):15–27. 21. Oades RD, Halliday GM. Ventral tegmental (A10) system: neurobiology. 1. Anatomy and connectivity. Brain Res Rev. 1987;12(2):117–165. 22. Fallon JH, Loughlin SE. Substantia nigra. In: Paxinos G, ed. The Rat Nervous System. 2nd ed. San Diego, CA: Academic Press; 1995:215–237. 23. Lewis DA, Sesack SR. Dopamine systems in the primate brain. In: Bloom FE, Bjo¨rklund A, Ho¨kfelt T, eds. Handbook of Chemical Neuroanatomy, The Primate Nervous System, Part I. Vol 13. Amsterdam: Elsevier Science Publishers; 1997:261–373. 24. Bjo¨rklund A, Dunnett SB. Dopamine neuron systems in the brain: an update. Trends Neurosci. 2007;30(5):194–202. 25. Peretti-Renucci R, Feuerstein C, Manier M, et al. Quantitative image analysis with densitometry for immunohistochemistry and autoradiography of receptor binding sites–methodological considerations. J Neurosci Res. 1991;28(4):583–600. 26. Maggio JE, Mantyh PW. Autoradiography of reversible ligands. In: Ariano MA, ed. Receptor Localization: Laboratory Methods and Procedures New York, NY: Wiley-Liss; 1998:17–30. 27. Chesselet MF. Localization of mRNAs encoding receptors with in situ hybridization histochemistry. In: Ariano MA, ed. Receptor Localization: Laboratory Methods and Procedures. New York, NY: Wiley-Liss; 1998:140–159. 28. Stornetta RL, Guyenet PG. Nonradioactive in situ hybridization in combination with tract-tracing. In: Zaborszky L, Wouterlood FG, Lanciego JL, eds. Neuroanatomical Tract-Tracing 3: Molecules, Neurons, Systems. New York, NY: Springer; 2006:237–262. 29. Mathiisen TM, Nagelhus EA, Jouleh B, et al. Postembedding immunogold cytochemistry of membrane molecules and amino acid transmitters in the central nervous system. In: Zaborszky L, Wouterlood FG, Lanciego JL, eds. Neuroanatomical TractTracing 3: Molecules, Neurons, Systems. New York, NY: Springer; 2006:72–108. 30. Sesack SR, Miner LAH, Omelchenko N. Pre-embedding immunoelectron microscopy: applications for studies of the nervous system. In: Zaborszky L, Wouterlood FG, Lanciego JL, eds. Neuroanatomical Tract-Tracing 3: Molecules, Neurons, Systems. New York, NY: Springer; 2006:6–71. 31. Surmeier DJ, Song W-J, Yan Z. Coordinated expression of dopamine receptors in neostriatal medium spiny neurons. J Neurosci. 1996;16(20):6579–6591. 32. Eberwine J, Belt B, Kacharmina JE, Miyashiro K. Analysis of subcellularly localized mRNAs using in situ hybridization, mRNA amplification, and expression profiling. Neurochem Res. 2002;27(10):1065–1077. 33. Saper CB. Magic peptides, magic antibodies: guidelines for appropriate controls for immunohistochemistry. J Comp Neurol. 2003;465(2):161–163. 34. Brey EM, Lalani Z, Johnston C, et al. Automated selection of DAB-labeled tissue for immunohistochemical quantification. J Histochem Cytochem. 2003;51(5):575–584. 35. Aizman O, Brismar H, Uhlen P, et al. Anatomical and physiological evidence for D1 and D2 dopamine receptor colocalization in neostriatal neurons. Nature Neurosci. 2000;3(3): 226–230.

2.2: FUNCTIONAL IMPLICATIONS OF DOPAMINE D2 RECEPTOR LOCALIZATION 36. Deng YP, Lei WL, Reiner A. Differential perikaryal localization in rats of D1 and D2 dopamine receptors on striatal projection neuron types identified by retrograde labeling. J Chem. Neuroanat. 2006;32(2-4):101–116. 37. Levey A, Hersch S, Rye D, et al. Localization of D1 and D2 dopamine receptors in brain with subtype-specific antibodies. Proc Natl Acad Sci USA. 1993;90(19):8861–8865. 38. Sesack SR, Aoki C, Pickel VM. Ultrastructural localization of D2 receptor-like immunoreactivity in midbrain dopamine neurons and their striatal targets. J Neurosci. 1994;14(1):88–106. 39. Somogyi P, Tama´s G, Lujan R, Buhl EH. Salient features of synaptic organisation in the cerebral cortex. Brain Res Rev. 1998;26(2-3):113–135. 40. Mercuri NB, Saiardi A, Bonci A, et al. Loss of autoreceptor function in dopaminergic neurons from dopamine D2 receptor deficient mice. Neuroscience. 1997;79(2):323–327. 41. Geisler S, Derst C, Veh RW, Zahm DS. Glutamatergic afferents of the ventral tegmental area in the rat. J Neurosci. 2007;27(21):5730–5743. 42. Bouthenet ML, Souil E, Martres MP, Sokoloff P, Giros B, Schwartz JC. Localization of dopamine D3 receptor mRNA in the rat brain using in situ hybridization histochemistry: comparison with dopamine D2 receptor mRNA. Brain Res. 1991;564(2):203–219. 43. Ariano MA, Fisher RS, Smyk-Randall E, Sibley DR, Levine MS. D2 dopamine receptor distribution in the rodent CNS using antipeptide antibodies. Brain Res. 1993;609(1-2):71–80. 44. Meador-Woodruff JH, Damask SP, Watson SJ Jr. Differential expression of autoreceptors in the ascending dopamine systems of the human brain. Proc Natl Acad Sci USA. 1994;91(17):8297–8301. 45. Choi WS, Machida CA, Ronnekleiv OK. Distribution of dopamine D1, D2, and D5 receptor mRNAs in the monkey brain: ribonuclease protection assay analysis. Mol Brain Res. 1995; 31(1-2):86–94. 46. Khan ZU, Gutierrez A, Martin R, Penafiel A, Rivera A, De La Calle A. Differential regional and cellular distribution of dopamine D2-like receptors: an immunocytochemical study of subtype-specific antibodies in rat and human brain. J Comp Neurol. 1998;402(3):353–371. 47. Pickel VM, Chan J, Nirenberg MJ. Region-specific targeting of dopamine D2-receptors and somatodendritic vesicular monoamine transporter 2 (VMAT2) within ventral tegmental area subdivisions. Synapse. 2002;45(2):113–124. 48. Tzschentke TM. Pharmacology and behavioral pharmacology of the mesocortical dopamine system. Prog Neurobiol. 2001;63(3):241–320. 49. Kita H, Kitai ST. Efferent projections of the subthalamic nucleus in the rat: light and electron microscopic analysis with the PHAL method. J Comp Neurol. 1987;260(3):435–452. 50. Omelchenko N, Sesack SR. Glutamate synaptic inputs to ventral tegmental area neurons in the rat derive primarily from subcortical sources. Neuroscience. 2007;146(3):1259–1274. 51. Bolam JP, Smith Y. The GABA and substance P input to dopaminergic neurones in the substantia nigra of the rat. Brain Res. 1990;529(1-2):57–78. 52. Kosinski CM, Standaert DG, Testa CM, Penney JB Jr, Young AB. Expression of metabotropic glutamate receptor 1 isoforms in the substantia nigra pars compacta of the rat. Neuroscience. 1998;86(3):783–798. 53. Chatha BT, Bernard V, Streit P, Bolam JP. Synaptic localization of ionotropic glutamate receptors in the rat substantia nigra. Neuroscience. 2000;101(4):1037–1051.

33

54. Rodriguez JJ, Doherty MD, Pickel VM. N-methyl-D-aspartate (NMDA) receptors in the ventral tegmental area: subcellular distribution and colocalization with 5-hydroxytryptamine(2A) receptors. J Neurosci Res. 2000;60(2):202–211. 55. Jones S, Kornblum JL, Kauer JA. Amphetamine blocks long-term synaptic depression in the ventral tegmental area. J Neurosci. 2000;20(15):5575–5580. 56. Thomas MJ, Malenka RC, Bonci A. Modulation of long-term depression by dopamine in the mesolimbic system. J Neurosci. 2000;20(15):5581–5586. 57. Sulzer D, Joyce MP, Lin L, et al. Dopamine neurons make glutamatergic synapses in vitro. J Neurosci. 1998;18(12):4588–4602. 58. Congar P, Bergevin A, Trudeau LE. D2 receptors inhibit the secretory process downstream from calcium influx in dopaminergic neurons: implication of K+ channels. J Neurophysiol. 2002;87(2):1046–1056. 59. Lavin A, Nogueira L, Lapish CC, Wightman RM, Phillips PE, Seamans JK. Mesocortical dopamine neurons operate in distinct temporal domains using multimodal signaling. J Neurosci. 2005;25(20):5013–5023. 60. Kawano M, Kawasaki A, Sakata-Haga H, et al. Particular subpopulations of midbrain and hypothalamic dopamine neurons express vesicular glutamate transporter 2 in the rat brain. J Comp Neurol. 2006;498(5):581–592. 61. Yamaguchi T, Sheen W, Morales M. Glutamatergic neurons are present in the rat ventral tegmental area. Eur J Neurosci. 2007;25(1):106–118. 62. Nair-Roberts RG, Chatelain-Badie SD, Benson E, White-Cooper H, Bolam JP, Ungless MA. Stereological estimates of dopaminergic, GABAergic and glutamatergic neurons in the ventral tegmental area, substantia nigra and retrorubral field in the rat. Neuroscience. 2008;152(4):1024–1031. 63. Fujiyama F, Kuramoto E, Okamoto K, et al. Presynaptic localization of an AMPA-type glutamate receptor in corticostriatal and thalamostriatal axon terminals. Eur J Neurosci. 2004;20(12):3322–3330. 64. Hur EE, Zaborszky L. Vglut2 afferents to the medial prefrontal and primary somatosensory cortices: a combined retrograde tracing in situ hybridization. J Comp Neurol. 2005; 483(3):351–373. 65. Dobi A, Morales M. Dopaminergic neurons in the rat ventral tegmental area (VTA) receive glutamatergic inputs from local glutamatergic neurons. Soc Neurosci Abstr. 2007:916–918. 66. Trudeau LE. Glutamate co-transmission as an emerging concept in monoamine neuron function. J Psychiatry Neurosci. 2004;29(4):296–310. 67. Lapish CC, Kroener S, Durstewitz D, Lavin A, Seamans JK. The ability of the mesocortical dopamine system to operate in distinct temporal modes. Psychopharmacology. 2007;191(3):609–625. 68. Goldman-Rakic PS, Leranth C, Williams SM, Mons N, Geffard M. Dopamine synaptic complex with pyramidal neurons in primate cerebral cortex. Proc Natl Acad Sci USA. 1989; 86(22):9015–9019. 69. Gasbarri A, Sulli A, Packard MG. The dopaminergic mesencephalic projections to the hippocampal formation in the rat. Prog Neuro-Psychopharm Biol Psychiatr. 1997;21(1):1–22. 70. Melchitzky DS, Erickson SL, Lewis DA. Dopamine innervation of the monkey mediodorsal thalamus: location of projection neurons and ultrastructural characteristics of axon terminals. Neuroscience. 2006;143(4):1021–1030. 71. Muller JF, Mascagni F, McDonald AJ. Dopaminergic innervation of pyramidal cells in the rat basolateral amygdala. Brain Struct Funct. 2009;213(3):275–288.

34

NEUROANATOMY

72. Smith Y, Kieval JZ. Anatomy of the dopamine system in the basal ganglia. Tr Neurosci. 2000;23(suppl 10):S28–S33. 73. Smith Y, Villalba R. Striatal and extrastriatal dopamine in the basal ganglia: an overview of its anatomical organization in normal and Parkinsonian brains. 2008;23:S534–S547. 74. Svenningsson P, Le Moine C. Dopamine D1/5 receptor stimulation induces c-fos expression in the subthalamic nucleus: possible involvement of local D5 receptors. Eur J Neurosci. 2002;15(1):133–142. 75. Gaspar P, Bloch B, Le Moine C. D1 and D2 receptor gene expression in the rat frontal cortex: cellular localization in different classes of efferent neurons. Eur J Neurosci. 1995; 7(5):1050–1063. 76. Meador-Woodruff JH, Damask SP, Wang JC, Haroutunian V, Davis KL, Watson SJ. Dopamine receptor mRNA expression in human striatum and neocortex. Neuropsychopharmacology. 1996;15(1):17–29. 77. Lidow MS, Wang F, Cao Y, Goldman-Rakic PS. Layer V neurons bear the majority of mRNAs encoding the five distinct dopamine receptor subtypes in the primate prefrontal cortex. Synapse. 1998;28(1):10–20. 78. de Almeida J, Palacios JM, Mengod G. Distribution of 5-HT and DA receptors in primate prefrontal cortex: implications for pathophysiology and treatment. Prog Brain Res. 2008;172:101–115. 79. Eliava M, Yilmazer-Hanke D, Asan E. Interrelations between monoaminergic afferents and corticotropin-releasing factorimmunoreactive neurons in the rat central amygdaloid nucleus: ultrastructural evidence for dopaminergic control of amygdaloid stress systems. Histochem Cell Biol. 2003;120(3):183–197. 80. Rieck RW, Ansari MS, Whetsell WO Jr, Deutch AY, Kessler RM. Distribution of dopamine D2-like receptors in the human thalamus: autoradiographic and PET studies. Neuropsychopharmacology. 2004;29(2):362–372. 81. Awenowicz PW, Porter LL. Local application of dopamine inhibits pyramidal tract neuron activity in the rodent motor cortex. J Neurophysiol. 2002;88(6):3439–3451. 82. Negyessy L, Goldman-Rakic PS. Subcellular localization of the dopamine D2 receptor and coexistence with the calcium-binding protein neuronal calcium sensor-1 in the primate prefrontal cortex. J Comp Neurol. 2005;488(4):464–475. 83. Pinto A, Sesack SR. Ultrastructural analysis of prefrontal cortical inputs to the rat amygdala: spatial relationships to presumed dopamine axons and D1 and D2 receptors. Brain Struct Funct. 2008;213(1-2):159–175. 84. Khan ZU, Koulen P, Rubinstein M, Grandy DK, GoldmanRakic PS. An astroglia-linked dopamine D2-receptor action in prefrontal cortex. Proc Natl Acad Sci USA. 2001;98(4):1964– 1969. 85. Agulhon C, Petravicz J, McMullen AB, et al. What is the role of astrocyte calcium in neurophysiology? Neuron. 2008;59(6): 932–946. 86. Kabbani N, Negyessy L, Lin R, Goldman-Rakic P, Levenson R. Interaction with neuronal calcium sensor NCS-1 mediates desensitization of the D2 dopamine receptor. J Neurosci. 2002;22(19):8476–8486. 87. Paspalas CD, Rakic P, Goldman-Rakic PS. Internalization of D2 dopamine receptors is clathrin-dependent and select to dendroaxonic appositions in primate prefrontal cortex. Eur J Neurosci. 2006;24(5):1395–1403. 88. Pickel VM, Chan J, Kearn CS, Mackie K. Targeting dopamine D2 and cannabinoid-1 (CB1) receptors in rat nucleus accumbens. J Comp Neurol. 2006;495(3):299–313.

89. Gulledge AT, Jaffe DB. Dopamine decreases the excitability of layer V pyramidal cells in the rat prefrontal cortex. J Neurosci. 1998;18(21):9139–9151. 90. Zheng P, Zhang XX, Bunney BS, Shi WX. Opposite modulation of cortical N-methyl-D-aspartate receptor-mediated responses by low and high concentrations of dopamine. Neuroscience. 1999;91(2):527–535. 91. Tseng KY, O’Donnell P. Dopamine-glutamate interactions controlling prefrontal cortical pyramidal cell excitability involve multiple signaling mechanisms. J Neurosci. 2004;24(22):5131– 5139. 92. Tseng KY, O’Donnell P. D2 dopamine receptors recruit a GABA component for their attenuation of excitatory synaptic transmission in the adult rat prefrontal cortex. Synapse. 2007;61(10):843–850. 93. Sun X, Zhao Y, Wolf ME. Dopamine receptor stimulation modulates AMPA receptor synaptic insertion in prefrontal cortex neurons. J Neurosci. 2005;25(32):7342–7351. 94. Kotecha SA, Oak JN, Jackson MF, et al. A D2 class dopamine receptor transactivates a receptor tyrosine kinase to inhibit NMDA receptor transmission. Neuron. 2002;35(6):1111–1122. 95. Kroner S, Rosenkranz JA, Grace AA, Barrionuevo G. Dopamine modulates excitability of basolateral amygdala neurons in vitro. J Neurophysiol. 2005;93(3):1598–1610. 96. Rosenkranz JA, Johnston D. State-dependent modulation of amygdala inputs by dopamine-induced enhancement of sodium currents in layer V entorhinal cortex. J Neurosci. 2007; 27(26):7054–7069. 97. O’Donnell P. Dopamine gating of forebrain neural ensembles. Eur J Neurosci. 2003;17(3):429–435. 98. Tseng KY, O’Donnell P. Post-pubertal emergence of prefrontal cortical up states induced by D1-NMDA co-activation. Cereb Cortex. 2005;15(1):49–57. 99. Chen L, Bohanick JD, Nishihara M, Seamans JK, Yang CR. Dopamine D1/5 receptor-mediated long-term potentiation of intrinsic excitability in rat prefrontal cortical neurons: Ca2+dependent intracellular signaling. J Neurophysiol. 2007;97(3): 2448–2464. 100. Otani S, Blond O, Desce JM, Crepel F. Dopamine facilitates long-term depression of glutamatergic transmission in rat prefrontal cortex. Neuroscience. 1998;85(3):669–676. 101. Gurden H, Takita M, Jay TM. Essential role of D1 but not D2 receptors in the NMDA receptor-dependent long-term potentiation at hippocampal-prefrontal cortex synapses in vivo. J Neurosci. 2000;20:RC106(22):1–5. 102. Gonzalez-Burgos G, Lewis DA. GABA neurons and the mechanisms of network oscillations: implications for understanding cortical dysfunction in schizophrenia. Schizophr Bull. 2008;34(5):944–961. 103. Sesack SR, Snyder CL, Lewis DA. Axon terminals immunolabeled for dopamine or tyrosine hydroxylase synapse on GABAimmunoreactive dendrites in rat and monkey cortex. J Comp Neurol. 1995;363(2):264–280. 104. Pinard CR, Muller JF, Mascagni F, McDonald AJ. Dopaminergic innervation of interneurons in the rat basolateral amygdala. Neuroscience. 2008;157(4):850–863. 105. Gulledge AT, Jaffe DB. Multiple effects of dopamine on layer V pyramidal cell excitability in rat prefrontal cortex. J Neurophysiol. 2001;86(2):586–595. 106. Tseng KY, Mallet N, Toreson KL, Le Moine C, Gonon F, O’Donnell P. Excitatory response of prefrontal cortical fastspiking interneurons to ventral tegmental area stimulation in vivo. Synapse. 2006;59(7):412–417.

2.2: FUNCTIONAL IMPLICATIONS OF DOPAMINE D2 RECEPTOR LOCALIZATION 107. Tseng KY, O’Donnell P. Dopamine modulation of prefrontal cortical interneurons changes during adolescence. Cereb Cortex. 2007;17(5):1235–1240. 108. Totterdell S, Smith AD. Convergence of hippocampal and dopaminergic input onto identified neurons in the nucleus accumbens of the rat. J Chem Neuroanat. 1989;2(5):285–298. 109. Sesack SR, Pickel VM. Prefrontal cortical efferents in the rat synapse on unlabeled neuronal targets of catecholamine terminals in the nucleus accumbens septi and on dopamine neurons in the ventral tegmental area. J Comp Neurol. 1992;320(2):145– 160. 110. Yao WD, Spealman RD, Zhang J. Dopaminergic signaling in dendritic spines. Biochem Pharmacol. 2008;75(11):2055–2069. 111. Henze DA, Gonzalez-Burgos GR, Urban NN, Lewis DA, Barrionuevo G. Dopamine increases excitability of pyramidal neurons in primate prefrontal cortex. J Neurophysiol. 2000;84(6):2799–2809. 112. Moss J, Bolam JP. A dopaminergic axon lattice in the striatum and its relationship with cortical and thalamic terminals. J Neurosci. 2008;28(44):11221–11230. 113. Boundy VA, Luedtke RR, Artymyshyn RP, Filtz TM, Molinoff PB. Development of polyclonal anti-D2 dopamine receptor antibodies using sequence-specific peptides. Mol Pharmacol. 1993;43(5):666–676. 114. Fisher RS, Levine MS, Sibley DR, Ariano MA. D2 dopamine receptor protein localization: golgi impregnation-gold toned and ultrastructural analysis of the rat neostriatum. J Neurosci Res. 1994;38(5):551–564. 115. Mengual E, Pickel VM. Ultrastructural immunocytochemical localization of the dopamine D2 receptor and tyrosine hydroxylase in the rat ventral pallidum. Synapse. 2002;43(3):151– 162. 116. Delle Donne KT, Sesack SR, Pickel VM. Ultrastructural immunocytochemical localization of neurotensin and the dopamine D2 receptor in the rat nucleus accumbens. J Comp Neurol. 1996;371(4):552–566. 117. Aubert I, Ghorayeb I, Normand E, Bloch B. Phenotypical characterization of the neurons expressing the D1 and D2 dopamine receptors in the monkey striatum. J Comp Neurol. 2000;418(1):22–32. 118. Alcantara AA, Chen V, Herring BE, Mendenhall JM, Berlanga ML. Localization of dopamine D2 receptors on cholinergic interneurons of the dorsal striatum and nucleus accumbens of the rat. Brain Res. 2003;986(1–2):22–29. 119. Gerfen CR, Engber TM, Mahan LC, et al. D1 and D2 dopamine receptor regulated gene expression of striatonigral and striatopallidal neurons. Science. 1990;250(4986):1429– 1432. 120. Le Moine C, Bloch B. D1 and D2 dopamine receptor gene expression in the rat striatum: sensitive cRNA probes demonstrate prominent segregation of D1 and D2 mRNAs in distinct neuronal populations of the dorsal and ventral striatum. J Comp Neurol. 1995;355(3):418–426. 121. Delle Donne KT, Sesack SR, Pickel VM. Ultrastructural immunocytochemical localization of the dopamine D2 receptor within GABAergic neurons of the rat striatum. Brain Res. 1997; 746(1-2):239–255. 122. Surmeier DJ, Eberwine J, Wilson CJ, Cao Y, Stefani A, Kitai ST. Dopamine receptor subtypes colocalize in rat striatonigral neurons. Proc Natl Acad Sci USA. 1992;89(21):10178–10182. 123. Surmeier DJ, Reiner A, Levine MS, Ariano MA. Are neostriatal dopamine receptors co-localized? Trends Neurosci. 1993;16(8):299–305.

35

124. Le Moine C, Bloch B. Expression of the D3 dopamine receptor in peptidergic neurons of the nucleus accumbens: comparison with the D1 and D2 dopamine receptors. Neuroscience. 1996;73(1):131–143. 125. Parent A, Sato F, Wu Y, Gauthier J, Levesque M, Parent M. Organization of the basal ganglia: the importance of axonal collateralization. Trends Neurosci. 2000;23(suppl 10): S20–S27. 126. Day M, Wang Z, Ding J, et al. Selective elimination of glutamatergic synapses on striatopallidal neurons in Parkinson disease models. Nat Neurosci. 2006;9(2):251–259. 127. Lee KW, Kim Y, Kim AM, Helmin K, Nairn AC, Greengard P. Cocaine-induced dendritic spine formation in D1 and D2 dopamine receptor-containing medium spiny neurons in nucleus accumbens. Proc Natl Acad Sci USA. 2006;103(9):3399–3404. 128. Wang Z, Kai L, Day M, et al. Dopaminergic control of corticostriatal long-term synaptic depression in medium spiny neurons is mediated by cholinergic interneurons. Neuron. 2006;50(3):443– 452. 129. Delle Donne KT, Chan J, Boudin H, Pelaprat D, Rostene W, Pickel VM. Electron microscopic dual labeling of high-affinity neurotensin and dopamine D2 receptors in the rat nucleus accumbens shell. Synapse. 2004;52(3):176–187. 130. Herna´ndez-Echeagaray E, Starling AJ, Cepeda C, Levine MS. Modulation of AMPA currents by D2 dopamine receptors in striatal medium-sized spiny neurons: are dendrites necessary? Eur J Neurosci. 2004;19(9):2455–2463. 131. Surmeier DJ, Ding J, Day M, Wang Z, Shen W. D1 and D2 dopamine-receptor modulation of striatal glutamatergic signaling in striatal medium spiny neurons. Trends Neurosci. 2007;30(5):228–235. 132. Smith AD, Bolam JP. The neural network of the basal ganglia as revealed by the study of synaptic connections of identified neurones. Trends Neurosci. 1990;13(7):259–265. 133. Bernard V, Somogyi P, Bolam JP. Cellular, subcellular, and subsynaptic distribution of AMPA-type glutamate receptor subunits in the neostriatum of the rat. J Neurosci. 1997;17(2):819– 833. 134. Guigoni C, Doudnikoff E, Li Q, Bloch B, Bezard E. Altered D(1) dopamine receptor trafficking in parkinsonian and dyskinetic non-human primates. Neurobiol Dis. 2007;26(2):452– 463. 135. Wang H, Pickel VM. Dopamine D2 receptors are present in prefrontal cortical afferents and their targets in patches of the rat caudate-putamen nucleus. J Comp Neurol. 2002;442(4): 392–404. 136. Seamans JK, Gorelova NA, Yang CR. Contributions of voltagegated Ca2+ channels in the proximal versus distal dendrites to synaptic integration in prefrontal cortical neurons. J Neurosci. 1997;17(15):5936–5948. 137. Ehlers MD, Heine M, Groc L, Lee MC, Choquet D. Diffusional trapping of GluR1 AMPA receptors by input-specific synaptic activity. Neuron. 2007;54(3):447–460. 138. Wolf ME, Mangiavacchi S, Sun X. Mechanisms by which dopamine receptors may influence synaptic plasticity. Ann NY Acad Sci. 2003;1003:241–249. 139. Lei W, Jiao Y, Del Mar N, Reiner A. Evidence for differential cortical input to direct pathway versus indirect pathway striatal projection neurons in rats. J Neurosci. 2004;24(38):8289–8299. 140. Cepeda C, Andre VM, Yamazaki I, Wu N, Kleiman-Weiner M, Levine MS. Differential electrophysiological properties of dopamine D1 and D2 receptor-containing striatal medium-sized spiny neurons. Eur J Neurosci. 2008;27(3):671–682.

36

NEUROANATOMY

141. Shen W, Flajolet M, Greengard P, Surmeier DJ. Dichotomous dopaminergic control of striatal synaptic plasticity. Science. 2008;321(5890):848–851. 142. Stephens B, Mueller AJ, Shering AF, et al. Evidence of a breakdown of corticostriatal connections in Parkinson’s disease. Neuroscience. 2005;132(3):741–754. 143. Ingham CA, Hood SH, Taggart P, Arbuthnott GW. Plasticity of synapses in the rat neostriatum after unilateral lesion of the nigrostriatal dopaminergic pathway. J Neurosci. 1998;18(12): 4732–4743. 144. Villalba RM, Lee H, Smith Y. Dopaminergic denervation and spine loss in the striatum of MPTP-treated monkeys. Exp Neurol. 2008;215(2):220–227. 145. Solis O, Limon DI, Flores-Hernandez J, Flores G. Alterations in dendritic morphology of the prefrontal cortical and striatum neurons in the unilateral 6-OHDA-rat model of Parkinson’s disease. Synapse. 2007;61(6):450–458. 146. Smythies J. Redox mechanisms at the glutamate synapse and their significance: a review. Eur J Pharmacol. 1999;370(1):1–7. 147. Kihara T, Shimohama S, Sawada H, et al. Protective effect of dopamine D2 agonists in cortical neurons via the phosphatidylinositol 3 kinase cascade. J Neurosci Res. 2002;70(3):274–282. 148. Wang HD, Deutch AY. Dopamine depletion of the prefrontal cortex induces dendritic spine loss: reversal by atypical antipsychotic drug treatment. Neuropsychopharmacology. 2008;33(6): 1276–1286. 149. Pierce JP, Lewin GR. An ultrastructural size principle. Neuroscience. 1994;58(3):441–446. 150. Sorra KE, Fiala JC, Harris KM. Critical assessment of the involvement of perforations, spinules, and spine branching in hippocampal synapse formation. J Comp Neurol. 1998;398(2): 225–240. 151. Meshul CK, Stallbaumer RK, Taylor B, Janowsky A. Haloperidol-induced morphological changes in striatum are associated with glutamate synapses. Brain Res. 1994; 648(2): 181–195. 152. Meshul CK, Janowsky A, Casey DE, Stallbaumer RK, Taylor B. Coadministration of haloperidol and SCH-23390 prevents the increase in ‘‘perforated’’ synapses due to either drug alone. Neuropsychopharmacology. 1992;7(4):285–293. 153. Robinson TE, Kolb B. Structural plasticity associated with exposure to drugs of abuse. Neuropharmacology. 2004;47(suppl 1): 33–46. 154. Bamford NS, Zhang H, Schmitz Y, et al. Heterosynaptic dopamine neurotransmission selects sets of corticostriatal terminals. Neuron. 2004;42(4):653–663. 155. Yang CR, Mogenson GJ. Dopamine enhances terminal excitability of hippocampal-accumbens neurons via D2 receptor: role of dopamine in presynaptic inhibition. J Neurosci. 1986;6(8):2470–2478. 156. O’Donnell P, Grace AA. Tonic D2-mediated attenuation of cortical excitation in nucleus accumbens neurons recorded in vitro. Brain Res. 1994;634(1):105–112. 157. Cepeda C, Hurst RS, Altemus KL, et al. Facilitated glutamatergic transmission in the striatum of D2 dopamine receptor-deficient mice. J Neurophysiol. 2001;85(2):659–670. 158. Rosenkranz JA, Grace AA. Dopamine attenuates prefrontal cortical suppression of sensory inputs to the basolateral amygdala of rats. J Neurosci. 2001;21(11):4090–4103. 159. Yin HH, Lovinger DM. Frequency-specific and D2 receptormediated inhibition of glutamate release by retrograde endocannabinoid signaling. Proc Natl Acad Sci USA. 2006; 103(21):8251–8256.

160. Koga E, Momiyama T. Presynaptic dopamine D2-like receptors inhibit excitatory transmission onto rat ventral tegmental dopaminergic neurones. J Physiol. 2000;523(pt 1):163–173. 161. Sokoloff P, Giros B, Martres MP, Bouthenet ML, Schwartz JC. Molecular cloning and characterization of a novel dopamine receptor (D3) as a target for neuroleptics. Nature. 1990;347(6289):146–151. 162. Le´vesque D, Martres MP, Diaz J, et al. A paradoxical regulation of the dopamine D3 receptor expression suggests the involvement of an anterograde factor from dopamine neurons. Proc Natl Acad Sci USA. 1995;92(5):1719–1723. 163. Levant B. The D3 dopamine receptor: neurobiology and potential clinical relevance. Pharmacol Rev. 1997;49(3):231–252. 164. Heidbreder CA, Gardner EL, Xi ZX, et al. The role of central dopamine D3 receptors in drug addiction: a review of pharmacological evidence. Brain Res Rev. 2005;49(1):77–105. 165. Ariano AA, Sibley DR. Dopamine receptor distribution in the rat CNS: elucidation using anti-peptide antisera directed against D1A and D3 subtypes. Brain Res. 1994;649(1-2):95–110. 166. Diaz J, Levesque D, Lammers CH, et al. Phenotypical characterization of neurons expressing the dopamine D3 receptor in the rat brain. Neuroscience. 1995;65(3):731–745. 167. Suzuki M, Hurd YL, Sokoloff P, Schwartz JC, Sedvall G. D3 dopamine receptor mRNA is widely expressed in the human brain. Brain Res. 1998;779(1-2):58–74. 168. Diaz J, Pilon C, Le Foll B, et al. Dopamine D3 receptors expressed by all mesencephalic dopamine neurons. J Neurosci. 2000;20(23):8677–8684. 169. Stanwood GD, Artymyshyn RP, Kung MP, Kung HF, Lucki I, McGonigle P. Quantitative autoradiographic mapping of rat brain dopamine D3 binding with [(125)I]7-OH-PIPAT: evidence for the presence of D3 receptors on dopaminergic and nondopaminergic cell bodies and terminals. J Pharmacol Exp Ther. 2000;295(3):1223–1231. 170. Koeltzow TE, Xu M, Cooper DC, et al. Alterations in dopamine release but not dopamine autoreceptor function in dopamine D3 receptor mutant mice. J Neurosci. 1998;18(6):2231–2238. 171. Le Foll B, Diaz J, Sokoloff P. Neuroadaptations to hyperdopaminergia in dopamine D3 receptor-deficient mice. Life Sci. 2005;76(11):1281–1296. 172. Van Tol HH, Bunzow JR, Guan HC, et al. Cloning of the gene for a human dopamine D4 receptor with high affinity for the antipsychotic clozapine. Nature. 1991;350(6319): 610–614. 173. Mrzljak L, Bergson C, Pappy M, Huff R, Levenson R, GoldmanRakic PS. Localization of dopamine D4 receptors in GABAergic neurons of the primate brain. Nature. 1996;381(6579): 245–248. 174. Avale ME, Falzone TL, Gelman DM, Low MJ, Grandy DK, Rubinstein M. The dopamine D4 receptor is essential for hyperactivity and impaired behavioral inhibition in a mouse model of attention deficit/hyperactivity disorder. Mol Psychiatry. 2004;9(7):718–726. 175. Ariano MA, Wang J, Noblett KL, Larson ER, Sibley DR. Cellular distribution of the rat D4 dopamine receptor protein in the CNS using anti-receptor antisera. Brain Res. 1997; 752(1-2):26–34. 176. We˛dzony K, Chocyk A, Mac´kowiak M, Fijał K, Czyrak A. Cortical localization of dopamine D4 receptors in the rat brain—immunocytochemical study. J Physiol Pharmacol. 2000;51(2):205–221. 177. Berger MA, Defagot MC, Villar MJ, Antonelli MC. D4 dopamine and metabotropic glutamate receptors in cerebral cortex

2.2: FUNCTIONAL IMPLICATIONS OF DOPAMINE D2 RECEPTOR LOCALIZATION

178.

179.

180.

181.

and striatum in rat brain. Neurochem Res. 2001;26(4): 345–352. Rivera A, Cuellar B, Giron FJ, Grandy DK, de la Calle A, Moratalla R. Dopamine D4 receptors are heterogeneously distributed in the striosomes/matrix compartments of the striatum. J Neurochem. 2002;80(2):219–229. Rivera A, Pen˜afiel A, Megı´as M, et al. Cellular localization and distribution of dopamine D(4) receptors in the rat cerebral cortex and their relationship with the cortical dopaminergic and noradrenergic nerve terminal networks. Neuroscience. 2008;155(3):997–1010. Svingos AL, Periasamy S, Pickel VM. Presynaptic dopamine D4 receptor localization in the rat nucleus accumbens shell. Synapse. 2000;36(3):222–232. Noain D, Avale ME, Wedemeyer C, Calvo D, Peper M, Rubinstein M. Identification of brain neurons expressing the dopamine D4 receptor gene using BAC transgenic mice. Eur J Neurosci. 2006;24(9):2429–2438.

37

182. Martina M, Bergeron R. D1 and D4 dopaminergic receptor interplay mediates coincident G protein-independent and dependent regulation of glutamate NMDA receptors in the lateral amygdala. J Neurochem. 2008; 106(6):2421–2435. 183. Wang X, Zhong P, Gu Z, Yan Z. Regulation of NMDA receptors by dopamine D4 signaling in prefrontal cortex. J Neurosci. 2003;23(30):9852–9861. 184. Onn SP, Wang XB, Lin M, Grace AA. Dopamine D1 and D4 receptor subtypes differentially modulate recurrent excitatory synapses in prefrontal cortical pyramidal neurons. Neuropsychopharmacology. 2006;31(2): 318–338. 185. Wang X, Zhong P, Yan Z. Dopamine D4 receptors modulate GABAergic signaling in pyramidal neurons of prefrontal cortex. J Neurosci. 2002;22(21):9185–9193. 186. Rubinstein M, Cepeda C, Hurst RS, et al. Dopamine D4 receptor-deficient mice display cortical hyperexcitability. J Neurosci. 2001;21(11):3756–3763.

2.3

Convergence of Limbic, Cognitive, and Motor Cortico -Striatal Circuits with Dopamine Pathways in Primate Brain SUZ ANN E N. HAB ER

INTRODUCTION Dopamine plays a central role in a wide variety of behaviors including reward, cognition, and motor control. Subpopulations of dopamine neurons have been associated with these different functions: the mesolimbic, mesocortical, and nigrostriatal pathways, respectively. Recently, all dopamine cell groups have been associated with the development of reward-based learning, leading to goal-directed behaviors. These behaviors require a complex interface between motivational drive, cognition, and action planning.1,2 The dopamine cells are an integral part of the basal ganglia. They send a massive output to the striatum, the main input structure of the basal ganglia. Moreover, this is a bidirectional pathway, with the dopamine cells receiving a major input from the striatum. Historically, the basal ganglia is best known for their motor functions, in large part because of the association between its neuropathology and neurodegenerative disorders affecting motor control. In particular, the degeneration of the substantia nigra pars compacta, is clearly linked to Parkinson’s disease. While a role for the basal ganglia in the control of movement is clear, our concept of basal ganglia function has dramatically changed in the past 20 years. It is now recognized to mediate the full range of behaviors leading to the development and execution of action plans, including the emotions, motivation, and cognition that drive them. Regions within each of the basal ganglia nuclei have been identified as serving these functions. The ventral striatum plays a key role in reward and reinforcement, central striatal areas are involved in executive functions, and dorsolateral regions are associated with sensorimotor control. Likewise, within the midbrain, the ventral tegmental area (VTA) is most closely linked to reward and reinforcement, while the 38

substantia nigra, pars compacta is linked to sensorimotor function. Thus, this set of subcortical nuclei works in tandem with cortex (particularly frontal cortex) via complex cortico-basal ganglia networks that are fundamentally linked to incentive-based learning and the development and execution of goal-directed behaviors. The basal ganglia are traditionally considered to process this information in parallel and segregated functional streams consisting of reward (limbic), associative (cognitive), and motor control circuits.3 Moreover, microcircuits within each region are thought to mediate different aspects of each function.4 However, while frontal cortex is indeed divided based on specific functions, expressed behaviors are the result of a combination of complex information processing that involves all of the frontal cortex. Indeed, appropriate responses to environmental stimuli require continual updating, learning to adjust behaviors according to new data. This necessitates coordination between limbic, cognitive, and motor systems to form smoothly executed, goal-directed behaviors. Parallel processing of functional information through different basal ganglia circuits does not address how information flows between circuits, thereby developing new behaviors (or actions) or adapting to those previously learned. While the anatomical pathways appear to be generally topographic from cortex through basal ganglia circuits, and while there are some physiological correlates to the functional domains of the striatum, a large body of growing evidence supports a duel processing system in which not only is information processed in parallel, but also integration occurs between functional circuits.5–9 This chapter will first briefly review the basic circuitry that underlies parallel processing; second, the anatomical basis for integration across different corticobasal ganglia circuits, with a particular emphasis on

2.3: CONVERGENCE OF LIMBIC, COGNITIVE, AND MOTOR CORTICO-STRIATAL CIRCUITS

dopamine; and finally, functional support for integrative processes. While the focus is on primate studies, key rodent experiments are also highlighted when primate data are unavailable.

PARALLEL PROCESSING Functional Organization of Frontal Cortex Frontal cortex is organized in a hierarchical manner and can be divided into functional regions10: the orbital (OFC) and anterior cingulate (ACC) prefrontal cortices are involved in emotions and motivation, the dorsal prefrontal cortex (DPFC) is involved in higher cognitive processes or executive functions, and the premotor and motor areas are involved in motor planning and the execution of those plans. The ACC is divided into ventral, or subgenual, ACC and dorsal ACC (dACC) areas. Medial orbital area 14 and the subgenual ACC cortex are collectively referred to as the ventral medial prefrontal cortex (vmPFC) and are particularly important in the expression of emotion.11,12 The OFC is involved in the development of reward-based learning, aversive, and goal-directed behaviors.13–17 This area receives input from multimodal sensory regions and is closely linked to the vmPFC.18,19 Lesions of the OFC, vmPFC, and dACC areas result in an inability to initiate and carry out goal-directed behaviors, and lead to socially inappropriate and impulsive behaviors.20–23 The DPFC is involved in working memory, set shifting, and strategic planning, often referred to as executive functions.24–28 Motor cortices are the most clearly defined areas of the frontal cortex. Caudal motor areas are highly microexcitable, closely timed to the execution of movement, and send a direct descending projection to spinal motor nuclei. Rostral motor areas are involved in sequence generation and motor learning. They are less microexcitable than the caudal motor areas but more so than the prefrontal cortex (PFC).29,30 Each of these frontal areas projects to specific striatal regions. However, in addition to the well-described topographic organization, they also follow non-topographic rules. Functional Projections Through the Basal Ganglia Together, the frontal regions that mediate reward, motivation, and affect regulation project primarily to the rostral striatum, including the n. accumbens, the medial caudate n., and the medial and ventral rostral putamen, collectively referred to as the ventral striatum. While the ventral striatum is similar to the

39

dorsal striatum in most respects, it also has some unique features. The ventral striatum contains a subterritory, called the shell, which has been shown in rodents to play a particularly important role in the circuitry underlying goal-directed behaviors, behavioral sensitization, and changes in affective states.31,32 Moreover, the ventral striatum alone receives a dense projection from the amygdala and from the hippocampus.33,34 The hippocampal projection is mostly limited to the shell, while the amygdala projects throughout a wider ventral striatal area. There are no clear histochemical boundaries between the ventral and dorsal striatum. Thus, the best way to define the ventral striatum is by its afferent projections, primarily the vmPFC, OFC, dACC, and the medial temporal lobe, particularly the amygdala.35 The vmPFC projects to the ventral medial striatum, including the shell, and extends along the medial edge of the dorsal ventral caudate n.9 The shell receives the densest innervation from medial areas 25, and 32 and from agranular insular cortex. The lateral orbital regions project to the central and lateral parts of the ventral striatum and extend into the central rostral caudate n. The dACC terminates in a wide medial striatal region, overlapping with inputs from both the OFC and vmPFC. Consistent with these inputs to the ventral striatum, physiological and imaging studies demonstrate the important role of this ventral striatal region in reward-based learning and motivation.36–38 The DPFC projects to the head of the caudate n. and to the putamen rostral to the anterior commissure. Caudal to the commissure, this projection is confined to the medial, central portion of the head of the caudate n., with few terminals in the central and caudal putamen.9,39 Different parts of the DPFC project with complex topography to different parts of the rostral caudate and putamen.40 Physiological, imaging, and lesion studies support the idea that these areas are involved in working memory and strategic planning processes, working together with the DPFC in mediating this function.41–43 Rostral premotor areas terminate in both the caudate and putamen, bridging the two with a continuous projection. Projections from caudal motor areas terminate almost entirely in the dorsolateral putamen, caudal to the anterior commissure. Few terminals are found rostral to the anterior commissure. Both caudal and rostral motor areas occupy much of the putamen caudal to the anterior commissure, a region that also receives overlapping projections from somatosensory cortex, resulting in a somatotopically organized sensorimotor area.44–46 In summary, projections from frontal cortex form a functional gradient of inputs from the ventromedial sector through the

40

NEUROANATOMY

dorsolateral striatum, with the medial and orbital PFCs terminating in the ventromedial part and the motor cortex terminating in the dorsolateral region. Like corticostriatal projection, thalamstriatal projections are organized in a general topographical manner such that interconnected and functionally associated thalamic and cortical regions terminate in the same general striatal region.47 The striatal projection to the pallidal complex and substantia nigra pars reticulata are also generally topographically organized, thus maintaining the functional organization of the striatum in these output nuclei.4,48–51 The ventral striatum terminates in the ventral pallidum and in the dorsal part of the midbrain. Terminals from the central striatum terminate more centrally in both the pallidum and the pars reticulata, while those from the sensorimotor areas of the striatum innervate the ventrolateral part of each pallidal segment and the ventrolateral substantia nigra.. Finally, the pallidum and pars reticulata project to the different basal ganglia output nuclei of the thalamus, the mediodorsal, and the ventral anterior and ventral lateral cell groups. The thalamic-cortical pathway is the last link in the circuit, and the outputs from the mediodorsal, ventral anterior, and ventral lateral thalamic n. are connected respectively to the collective limbic areas, the associative control areas, and the motor control areas.4,52–55 Thus, the organization of connections through the corticobasal ganglia cortical network preserves a general functional topography within each structure, from the

cortex through the striatum, from the striatum to the pallidum/pars reticulata, from these output structures to the thalamus, and finally, back to the cortex (Fig. 2.3.1). This organization has led to the concept that each functionally identified cortical region drives (and is driven by) a specific basal ganglia loop or circuit, leading, in turn, to the idea of parallel processing of cortical information through segregated basal ganglia circuits.3 This concept focuses on the role of the basal ganglia in the selection and implementation of an appropriate motor response while inhibiting unwanted ones.56 The model assumes, however, that the behavior has been learned and that the role of the basal ganglia is to carry out a coordinated action. We now know that the cortico-basal ganglia network is critical in mediating the learning process to adapt and to accommodate past experiences to modify behavioral responses.41,57–60 This requires some communication across circuits.

INTEGRATIVE PATHWAYS Growing evidence has identified possible anatomical substrates through which transfer of information can occur across functional domains.5–9 Integration between different aspects of reward processing, as well as interaction with cognitive and motor control regions, likely occur at several stations throughout the system. For example, as indicated above, while there is a general topographic organization to the dense (or focal)

DPFC Mo tor

CC C/A

OF

ASSOCIATIVE

MOTOR

Thalamus

LIMBIC

GP/SNr

FIGURE 2.3.1. Schematic illustrating parallel circuits through corticobasal ganglia pathways. Corresponding shaded striatal and cortical areas demonstrate topographic projections; white, limbic circuit; light gray, associative circuit; dark gray, motor control circuit.

2.3: CONVERGENCE OF LIMBIC, COGNITIVE, AND MOTOR CORTICO-STRIATAL CIRCUITS

corticostriatal terminal fields, this projection system also has non-topographic rules. Focal projections from different functional cortical areas also converge in specific striatal areas. These areas of convergence create nodal points of integration embedded within a generally parallel system. Such an arrangement may set the stage for a differential impact on midbrain dopamine cells during learning. In this chapter, we emphasize the role of dopamine in this transfer through its anatomical relationships to the corticostriatal network. First, however, we review the non-topographical aspects of the corticostriatal projection system. Corticostriatal Projections The ventral striatum, the area that receives input from the vmPFC, dACC, and OFC, is concentrated in the rostral striatum. Collectively, the terminal fields from these cortical areas occupy approximately 22% of the striatum. As noted above, terminal fields from these cortical areas are concentrated in different striatal regions. However, they also converge extensively.9

41

The areas in which convergence occurs may be particularly critical for the coordination of different aspects of affect regulation (Fig. 2.3.2a-b ). Projection fields from the DPFC terminate from the rostral pole and continue throughout much of the body of the caudate n. and medial putamen. However, while focal corticostriatal projections from different limbic and cognitive regions generally occupy separate positions in the striatum, they also converge at specific locations, primarily at the rostral levels (Fig. 2.3.2b). Here, terminals from the DPFC partially converge with those from both the dACC and OFC. In fact, projections from all PFC areas occupy a central region, with each cortical projection extending into nonoverlapping zones.9 Convergence is less prominent caudally, with almost complete separation of the dense terminals from the DPFC and dACC/OFC/vmPFC just rostral to the anterior commissure. This pattern of PFC projection fields implies a central role, particularly for rostral striatal subregions, in synchronizing different aspects of reward and learning for long-term strategic planning and habit formation.

FIGURE 2.3.2. Schematics demonstrating convergence of corticostriatal focal projections from different limbic, associative, and motor areas: (a,b) convergence between projections from different prefrontal regions; (c) convergence between prefrontal regions and motor control areas. DPFC, dorsal prefrontal cortex; OFC, orbital prefrontal cortex; vmPFC, ventral medial prefrontal cortex. DPFC=dorsal prefrontal cortex; OFC=orbital prefrontal cortex; vmPFC=ventral, medial prefrontal cortex.

42

NEUROANATOMY

Just rostral to and at the level of the anterior commissure, convergence occurs between terminals from the DPFC and premotor regions.61 Interestingly, at more anterior levels, these projections remain relatively segregated. This is a place of prominent convergence between focal projections from the41 DPFC and those from the OFC/ACC/vmPFC. Figure 2.3.2c is a schematic of a coronal section just anterior to the anterior commissure demonstrating the relatively little convergence with limbic input but an interface with rostral motor control areas. Importantly, there are few convergent terminals between afferent projections from limbic and motor control regions. Projections from DPFC are therefore in a pivotal position in the striatum, converging at one level with inputs from areas associated with motivation and reward and, at a more posterior level, with those from cortical areas associated with action planning. Convergence between terminals from limbic and cognitive areas rostrally, and from cognitive and premotor motor regions more caudally, provides a possible neural substrate for executive control over the development of incentive-based actions. Taken together, the frontostriatal network therefore constitutes a dual system comprising both clearly segregated connections and subregions that contain convergent pathways derived from functionally discrete cortical areas. This dual projection system is further supported using probabilistic tractography methods in humans.8 It provides an anatomical substrate for a recent finding in rodents in which cross-encoding cortical information influenced the future firing of medium spiny neurons.62 The nodal points of convergence from different cortical regions may therefore constitute zones for dynamic restructuring of neural ensembles fundamental to learning. These subregions are in a position to send a more functionally integrated input to the dopamine cells compared to the majority of the striatum. Moreover, dopamine input to these subregions is likely to have a different impact on information flow through the basal ganglia circuits. The Midbrain Dopamine System The striatal incentive-related learning process is thought to originate, in part, from the midbrain dopamine cells, which signal reward prediction error or reward saliency.1,63,64 However, the latency between the presentation of the stimuli and the activity of the dopamine cells is too short to reflect the higher cortical processing necessary for linking a stimulus with its rewarding properties.65 This is consistent with the fact that in behavioral studies, animals have already been trained to link the response to the reward. Indeed,

in studies that associate reward prediction error with the dopamine neurons, animals are overtrained to recognize the reward. In other words, the animals first must learn to associate the brown, wrinkled, sticky substance with the sweet taste of a raisin. A critical issue, therefore, is, how do the dopamine cells receive information consolidating this association? The largest forebrain input to the dopamine neurons is from the striatum, and the largest input to the striatum is from cortex. Collectively, the PFC inputs to the striatum are in a position to modulate the striatal response to different aspects of reward saliency and value. Thus, although the short-latency burst firing activity of dopamine that signals the immediate reinforcement is likely to be triggered from brainstem nuclei, the cortico-striato-midbrain pathway is in a position to ‘‘train’’ dopamine cells to distinguish rewards in order to calculate error prediction. The nodal points of convergence between different functional cortical pathways within the striatum may be critical in this initial role and play a particularly important role in the temporal training of dopamine cells, placing these cells in a position to respond with a short-latency signal derived from incoming sensory systems.65 Over time, the fast burst firing activity of the dopamine cells is quickly activated by the brainstem as incoming stimuli are perceived. This, in turn, impacts the striatum and can influence progressively more dorsal regions during learning and the development of habit formation.66–68

The organization of dopamine neurons Anatomically, the midbrain dopamine neurons are not clearly defined within the mesolimbic, mesocortical, and nigrostriatal categories. In rodents, the midbrain dopamine neurons are generally divided into the substantia nigra pars compacta (SNc), the VTA, and the retrorubral cell groups.69 In primates, the SNc is further divided into three groups: a dorsal group (the a group); a main densocellular region (the ß group); and a ventral group (the g group), or cell columns.70,71 The dorsal group is oriented horizontally and extends dorsolaterally, circumventing the ventral and lateral superior cerebellar peduncle and the red nucleus. These cells merge with the immediately adjacent dopamine cell groups of the VTA to form a continuous mediodorsal band of cells. Calbindin, a calcium binding protein (CaBP), marks both the VTA and the dorsal SNc. In contrast, the ventral cell groups (the densocellular group and the cell columns) are calbindin negative and, unlike the dorsal tier, have high expression levels for the dopamine

2.3: CONVERGENCE OF LIMBIC, COGNITIVE, AND MOTOR CORTICO-STRIATAL CIRCUITS

Red N.

VTA

SNc SNr Dorsal tier Ventral tier

FIGURE 2.3.3. Schematic illustrating the organization of the midbrain dopamine neurons into the dorsal and ventral tiers. SNc, substantia nigra pars compacta; SNr, substantia nigra pars reticulata; VTA, ventral tegmental area.

transporter and for the D2 receptor mRNAs.71,72 Thus, the midbrain cells are divided into a dorsal tier that includes the VTA and the dorsal SNc and a ventral tier that includes the densocellular group and cell columns (Fig. 2.3.3). Afferent projections Input to the midbrain dopamine neurons is primarily from the striatum, from both the external segment of the globus pallidus and the ventral pallidum, and from the brainstem (for review, see73). Descending projections from the central nucleus of the amygdala also terminate in a wide mediolateral region but are limited primarily to the dorsal tier cells. In addition, there are projections to the dorsal tier from the bed nucleus of the stria terminalis and from the sublenticular substantia innominata that travel together with those from the amygdala.74 While the dopamine neurons receive input from these several sources, perhaps the most massive projection is from the striatum. Striatal projections terminate on both the dorsal and ventral

VTA

SNc

43

tiers in addition to the pars reticulata. This afferent projection is organized with an inverse ventral/dorsal topography. The ventral striatum projects widely to the dorsal tier and much of the dorsal part of the densocellular pars compacta cells. This ventral striatal terminal field extends laterally to include a large mediolateral region. Descending projections from the extended amygdala also terminate in a wide mediolateral region but primarily in the dorsal tier. Therefore, the dorsal tier receives a massive limbic input through an indirect projection from the OFC/dACC/vmPFC (via the striatum) and a direct projection from the extended amygdala. The central striatum, which receives input from the DPFC, projects extensively to the central and ventral parts of the densocellular region, extending into the cell columns and surrounding pars reticulata. Finally, the dorsolateral striatal projection is concentrated in the ventral and lateral parts of the substantia nigra. Unlike the widespread terminal fields of the ventral and central striatum, the distribution of efferent fibers from the dorsolateral striatum is more restricted and terminates primarily in the pars reticulata. However, their terminal fields do project to the cell columns of dopamine neurons that penetrate deep into the pars reticulata. Thus, in addition to the inverse dorsoventral topographic organization to the striatonigral projection, there is an important difference in the extent of the projection fields from the functional striatal domains. Projections from regions receiving PFC inputs have wide projection fields throughout the midbrain dopamine cells, while those from motor control areas have a relatively limited projection field (Fig. 2.3.4a).

Efferent projections Like the descending striatonigral pathway, the ascending nigrostriatal projection exhibits an inverse dorsoventral topographic arrangement. Here, there is

SNr CP

a

b

c

FIGURE 2.3.4. Schematic of the substantia nigra showing the combined distribution of striatonigral terminal fields (a, c) and nigrostriatal cells (b, c) associated with different functional regions of the striatum. Light gray, inputs and outputs from the limbic striatum; medium gray, inputs and outputs from the associative striatum; dark gray, inputs and outputs from the motor striatum. CP, cerebral peduncles; SNc, substantia nigra pars compacta; SNr, substantia nigra pars reticulata; VTA, ventral tegmental area.

44

NEUROANATOMY

also a mediolateral topographic organization. Thus, the dorsal and medial dopamine cells project to the ventral and medial parts of the striatum, while the ventral and lateral cells project to the dorsal and lateral parts of the striatum.5,48,75,76 Moreover, as with the striatonigral projection, the proportional distribution of cells that project to different functional domains of the striatum differs (Fig. 2.3.4b). The shell region of the ventral striatum receives the most limited midbrain input, primarily derived from the VTA. The rest of the ventral striatum receives input primarily from the dorsal tier, including the retrorubral cell group, and from the medial and dorsal regions of the densocellular group. The central part of the striatum, which receives input from the DPFC, also receives input from the central part of the densocellular region of the dopamine cells. In contrast, the dorsolateral part of the striatum receives input from a wide range of dopamine cells derived from the ventral tier, including both the densocellular and cell columns groups (Fig. 2.3.4b). When considered separately, each limb of the system creates a loose topographic organization. The VTA and medial substantia nigra are associated with the limbic regions, the central substantia nigra with associative regions, and the lateral and ventral substantia nigra are related to the motor control striatal regions (Fig. 2.3.4c). However, the fact that the descending and ascending limb of each functional striatonigral and nigrostriatal pathways differs in its proportional projections significantly alters the relationship of different functional striatal areas with the midbrain. The ventral striatum receives a relatively limited midbrain input but projects to a large region, which includes dorsal and ventral tiers and the dorsal pars reticulata. In contrast, the dorsolateral striatum receives input from a wide range of dopamine cells but projects to a limited region (Fig. 2.3.4c).5 The striato-nigro-striatal projection system The proportional differences between inputs and outputs of the dopamine neurons, coupled with their topography, result in complex interweaving of functional pathways. For each striatal region, the afferent and efferent striato-nigro-striatal projection system contains three components in the midbrain. There is a reciprocal connection that is flanked by two nonreciprocal connections. The reciprocal component contains cells that project to a specific striatal area. These cells are embedded within terminals from that same striatal area. Dorsal to this region lies a group of cells that project to the same striatal region but do not lie within its reciprocal terminal field. In other words, these cells receive a striatal projection from a region to which they do not project.

Finally, ventral to the reciprocal component are efferent terminals. However, there are no cells embedded in these terminals that project to that same specific striatal region. The cells that are located in this terminal field project to a different striatal area. These three components for each striato-nigro-striatal projection system occupy different positions within the midbrain. The ventral striatum system lies dorsomedially, the dorsolateral striatum system lies ventrolaterally, and the central striatal system is positioned between the two. Moreover, as indicated above, each functional region differs in its proportional projections that significantly alter their relationship to each other. The ventral striatum receives a limited midbrain input but projects to a large region. In contrast, the dorsolateral striatum receives a wide input but projects to a limited region. In other words, the ventral striatum influences a wide range of dopamine neurons but is itself influenced by a relatively limited group of dopamine cells. On the other hand, the dorsolateral striatum influences a limited midbrain region but is affected by a relatively large midbrain region. Thus, the size and position of the afferent and efferent connections for each system, together with the arrangement into three components, allow information from the limbic system to reach the motor system through a series of connections5 (see Fig. 2.3.5). The ventral striatum receives input from limbic regions and projects to the dorsal tier. The dorsal tier projects back to the ventral striatum. However, the ventral striatum efferent projection to the midbrain extends beyond the tight ventral striatal/ dorsal tier/ventral striatal circuit, terminating lateral and ventral to the dorsal tier. This area of terminal projection does not project back to the ventral striatum. Rather, cells in this region project more dorsally, into the striatal area that receives input from the DPFC. Through this connection, the same cortical information that influences the dorsal tier through the ventral striatum also modulates the densocellular region that projects to the central striatum. This central striatal region is reciprocally connected to the densocellular region. But it also projects to the ventral densocellular area and into the cell columns. Thus, projections from the DPFC, via the striatum, are in a position to influence cells that project to motor control areas of the striatum. The dorsolateral striatum is reciprocally connected to the ventral densocellular region and cell columns. The confined distribution of efferent dorsolateral striatal fibers limits the influence of the motor striatum to a relatively small region involving the cell columns and the pars reticulata. Taken together, the interface between different striatal

2.3: CONVERGENCE OF LIMBIC, COGNITIVE, AND MOTOR CORTICO-STRIATAL CIRCUITS

45

DPFC C /AC OFC

Mo to r Y Y O Y

O O

O

X O

Y

Y

X

X

O O X

Thalamus

Y

Y O

X

O X

X X

X X

X

GP/SNr

O O

X

VTA

X

SNc SNr Dorsal tier Ventral tier

FIGURE 2.3.5. Schematic illustrating both dual parallel and integrative processing through corticobasal ganglia pathways. Corresponding shaded striatal and cortical areas demonstrate topographic of projections. X marks substriatal regions where convergence between terminals from limbic cortical areas occurs; O, convergence between terminals from limbic and cognitive cortical areas; Y, convergence between terminals from cognitive and motor control cortical areas occurs. Arrows connecting the striatum and substantia nigra illustrate how the ventral striatum can influence the dorsal striatum through the midbrain dopamine cells. The connections between integrated areas also enter the parallel processing system, back to cortex, as indicated by the arrows connecting the striatum via the pallidum and thalamus. DPFC, dorsolateral prefrontal cortex; GP/SNr, globus pallidus/substantia nigra pars reticulata; OFC/ACC, orbital prefrontal/anterior cingulate cortex; SNc, substantia nigra pars compacta; SNr, substantia nigra pars reticulata; VTA, ventral tegmental area.

regions via the midbrain dopamine cells is organized in an ascending spiral interconnecting different functional regions of the striatum. This creates a feedforward organization (Fig. 2.3.5). Through this spiral of inputs and outputs between the striatum and midbrain dopamine neurons, information can be channeled from the shell and ventral striatum, through the central striatum, and to the dorsolateral striatum. In this way, information can flow from limbic to cognitive to motor circuits, providing a mechanism by which motivation and cognition can influence motor decisionmaking processes and appropriate responses to environmental cues. Functional Considerations A key component in developing appropriate goaldirected behaviors is the ability to first correctly evaluate different aspects of reward, including value versus risk and predictability, and inhibit maladaptive choices,

based on previous experience. These calculations rely on integration of different aspects of reward processing and cognition to develop and execute appropriate action plans. While parallel networks that mediate different functions are critical to maintaining coordinated behaviors, cross-talk between functional circuits during learning is critical. Indeed, reward and associative functions are not clearly and completely separated within the striatum. Consistent with human imaging studies, reward-responsive neurons are not restricted to the ventral striatum, but rather are found throughout the striatum. Moreover, cells responding in working memory tasks are often found also in the ventral striatum.37,43,77–79 As described above, embedded within limbic, associative, and motor control striatal territories are subregions containing convergent terminals between different reward-processing cortical areas, between these projections and those from the DPFC, and between the DPFC and rostral motor control areas.

46

NEUROANATOMY

Given that a single corticostriatal axon can innervate 14% of the striatum80 and that terminals from different cortical areas synapse on a subpopulation of interneurons that are important for integrating information across functions,81 these nodes of converging terminals may represent ‘‘hot spots’’ that may be particularly sensitive to synchronizing information across functional areas to impact on long-term strategic planning, and habit formation.62 Indeed, cells in the dorsal striatum are progressively recruited during different types of learning, from simple motor tasks to drug self-administration.41,66,82,83 Convergent fibers from cortex within the ventral striatum, taken together with hippocampal and amygdalo-striatal projections, place the ventral striatum in a key entry port for processing emotional and motivational information that, in turn, drives basal ganglia action output. The ventral, reward-based striatal region and the associative, central striatal region can impact on motor output circuits, not only through convergent terminal fields within the striatum, but also through the striato-nigro-striatal pathways. One can hypothesize that initially the nodal points of interface between the reward and associative circuits, for example, send a coordinated signal to dopamine cells. This pathway is in a pivotal position for temporal training dopamine cells. In turn, these nodal points may be further reinforced through the burst firing activity of the nigrostriatal pathway, thus transferring that impact back to the striatum (Fig. 2.3.5). Moreover, since the midbrain dopamine neurons project to a wider dorsal striatal region, information is transferred to other functional regions during learning and habit formation.66,67 This signal then enters the parallel system and, via the pallidum and thalamus, impacts on frontal cortex (Fig. 2.3.5). Indeed, when the striato-nigro-striatal circuit is interrupted, information transfer from Pavlovian to instrumental learning does not take place.84. Parallel circuits and integrative circuits must work together, allowing the coordinated behaviors to be maintained and focused (via parallel networks), but also to be modified and changed according to the appropriate external and internal stimuli (via integrative networks) (Fig. 2.3.5). Both the ability to maintain focus in the execution of specific behaviors and the ability to adapt appropriately to external and internal cues are key deficits in basal ganglia diseases that affect these aspects of motor control, cognition, and motivation. Within each interconnected corticobasal ganglia loop, there are subregions that cross functional domains. Their locations (within the striatum or midbrain) are likely to impact differentially on how the dopamine neurons mediate learning and the

development of action plans. Dopamine neurons are in a position, therefore, not only to impact on the striatum during learning, but also to be modulated by it during the development of learning and habit formation.66,82,84 REFERENCES 1. Schultz W. Getting formal with dopamine and reward. Neuron. 2002;36:241–263. 2. Matsumoto N, Hanakawa T, Maki S, Graybiel AM, Kimura M. Nigrostriatal dopamine system in learning to perform sequential motor tasks in a predictive manner. J Neurophysiol. 1999; 82:978–998. 3. Alexander GE, Crutcher MD. Functional architecture of basal ganglia circuits: neural substrates of parallel processing. Trends Neurosci. 1990;13:266–271. 4. Middleton FA, Strick PL. Basal-ganglia ‘projections’ to the prefrontal cortex of the primate. Cereb Cortex. 2002;12:926–935. 5. Haber SN, Fudge JL, McFarland NR. Striatonigrostriatal pathways in primates form an ascending spiral from the shell to the dorsolateral striatum. J Neurosci. 2000;20:2369–2382. 6. McFarland NR, Haber SN. Thalamic relay nuclei of the basal ganglia form both reciprocal and nonreciprocal cortical connections, linking multiple frontal cortical areas. J Neurosci. 2002;22:8117–8132. 7. Percheron G, Filion M. Parallel processing in the basal ganglia: up to a point. Trends Neurosci. 1991;14:55–59. 8. Draganski B, Kherif F, Kloppel S, Cook PA, Alexander DC, Parker GJ, Deichmann R, Ashburner J, Frackowiak RS. Evidence for segregated and integrative connectivity patterns in the human basal ganglia. J Neurosci. 2008;28:7143–7152. 9. Haber SN, Kim KS, Mailly P, Calzavara R. Reward-related cortical inputs define a large striatal region in primates that interface with associative cortical inputs, providing a substrate for incentive-based learning. J Neurosci. 2006;26:8368–8376. 10. Fuster JM. The prefrontal cortex–an update: time is of the essence. Neuron. 2001;30:319–333. 11. Mayberg HS, Liotti M, Brannan SK, McGinnis S, Mahurin RK, Jerabek PA, Silva JA, Tekell JL, Martin CC, Lancaster JL, Fox PT. Reciprocal limbic-cortical function and negative mood: converging PET findings in depression and normal sadness. Am J Psychiatry. 1999;156:675–682. 12. Milad MR, Quinn BT, Pitman RK, Orr SP, Fischl B, Rauch SL. Thickness of ventromedial prefrontal cortex in humans is correlated with extinction memory. Proc Natl Acad Sci USA. 2005;102:10706–10711. 13. Hikosaka K, Watanabe M. Delay activity of orbital and lateral prefrontal neurons of the monkey varying with different rewards. Cereb Cortex. 2000;10:263–271. 14. Schultz W, Tremblay L, Hollerman JR. Reward processing in primate orbitofrontal cortex and basal ganglia. Cereb Cortex. 2000;10:272–284. 15. Padoa-Schioppa C, Assad JA. Neurons in the orbitofrontal cortex encode economic value. Nature. 2006;441:223–226. 16. Kringelbach ML, Rolls ET. The functional neuroanatomy of the human orbitofrontal cortex: evidence from neuroimaging and neuropsychology. Prog Neurobiol. 2004;72:341–372. 17. O’Doherty J, Kringelbach ML, Rolls ET, Hornak J, Andrews C. Abstract reward and punishment representations in the human orbitofrontal cortex. Nat Neurosci. 2001;4:95–102.

2.3: CONVERGENCE OF LIMBIC, COGNITIVE, AND MOTOR CORTICO-STRIATAL CIRCUITS 18. Barbas H. Architecture and cortical connections of the prefrontal cortex in the rhesus monkey. In: Chauvel P, Delgado-Escueta AV, eds. Advances in Neurology. New York, NY: Raven Press; 1992:91–115. 19. Price JL, Carmichael ST, Drevets WC. Networks related to the orbital and medial prefrontal cortex; a substrate for emotional behavior? Prog Brain Res. 1996;107:523–536. 20. Butter CM, Snyder DR. Alterations in aversive and aggressive behaviors following orbital frontal lesions in rhesus monkeys. Acta Neurobiol Exp. 1972;32:525–565. 21. Fuster JM. Lesion studies. In: The Prefrontal Cortex: Anatomy, Physiology, and Neuropsychology of the Frontal Lobe. 2nd ed. New York, NY: Raven Press; 1989:51–82. 22. Milad MR, Rauch SL. The role of the orbitofrontal cortex in anxiety disorders. Ann NY Acad Sci. 2007;1121:546–561. 23. O’Doherty J, Critchley H, Deichmann R, Dolan RJ. Dissociating valence of outcome from behavioral control in human orbital and ventral prefrontal cortices. J Neurosci. 2003;23:7931–7939. 24. Smith EE, Jonides J. Working memory: a view from neuroimaging. Cogn Psychol. 1997;33:5–42. 25. Passingham D, Sakai K. The prefrontal cortex and working memory: physiology and brain imaging. Curr Opin Neurobiol. 2004;14:163–168. 26. Blumenfeld RS, Ranganath C. Dorsolateral prefrontal cortex promotes long-term memory formation through its role in working memory organization. J Neurosci. 2006;26: 916–925. 27. Fuster JM. Prefrontal neurons in networks of executive memory. Brain Res Bull. 2000;52:331–336. 28. Goldman-Rakic PS. The prefrontal landscape: implications of functional architecture for understanding human mentation and the central executive. Philos Trans R Soc Lond-Series B: Biol Sci. 1996;351:1445–1453. 29. Mushiake H, Inase M, Tanji J. Neuronal activity in the primate premotor, supplementary, and precentral motor cortex during visually guided and internally determined sequential movements. J Neurophysiol. 1991;66:705–718. 30. Tanji J, Mushiake H. Comparison of neuronal activity in the supplementary motor area and primary motor cortex. Brain Res Cogn Brain Res. 1996;3:143–150. 31. Ito R, Robbins TW, Everitt BJ. Differential control over cocaineseeking behavior by nucleus accumbens core and shell. Nat Neurosci. 2004;7:389–397. 32. Carlezon WA, Wise RA. Rewarding actions of phencyclidine and related drugs in nucleus accumbens shell and frontal cortex. J Neurosci. 1996;16:3112–3122. 33. Fudge JL, Kunishio K, Walsh C, Richard D, Haber SN. Amygdaloid projections to ventromedial striatal subterritories in the primate. Neuroscience. 2002;110:257–275. 34. Friedman DP, Aggleton JP, Saunders RC. Comparison of hippocampal, amygdala, and perirhinal projections to the nucleus accumbens: combined anterograde and retrograde tracing study in the macaque brain. J Comp Neurol. 2002; 450:345–365. 35. Haber SN, McFarland NR. The concept of the ventral striatum in nonhuman primates. In: McGinty JF, ed. Advancing from the Ventral Striatum to the Extended Amygdala. New York, NY: New York Academy of Sciences; 1999:33–48. 36. Tremblay L, Hollerman JR, Schultz W. Modifications of reward expectation-related neuronal activity during learning in primate striatum. J Neurophysiol. 1998;80:964–977. 37. Hassani OK, Cromwell HC, Schultz W. Influence of expectation of different rewards on behavior-related neuronal activity in the striatum. J Neurophysiol. 2001;85:2477–2489.

47

38. Knutson B, Adams CM, Fong GW, Hommer D. Anticipation of increasing monetary reward selectively recruits nucleus accumbens. J Neurosci. 2001;21:RC159. 39. Selemon LD, Goldman-Rakic PS. Longitudinal topography and interdigitation of corticostriatal projections in the rhesus monkey. J Neurosci. 1985;5:776–794. 40. Calzavara R, Mailly P, Haber SN. Relationship between the corticostriatal terminals from areas 9 and 46, and those from area 8A, dorsal and rostral premotor cortex and area 24c: an anatomical substrate for cognition to action. Eur J Neurosci 2007; 26:2005–2024. 41. Pasupathy A, Miller EK. Different time courses of learning-related activity in the prefrontal cortex and striatum. Nature. 2005;433:873–876. 42. Battig K, Rosvold HE, Mishkin M. Comparison of the effect of frontal and caudate lesions on delayed response and alternation in monkeys. J Comp Physiol Psychol. 1960;53: 400–404. 43. Levy R, Friedman HR, Davachi L, Goldman-Rakic PS. Differential activation of the caudate nucleus in primates performing spatial and nonspatial working memory tasks. J Neurosci. 1997;17: 3870–3882. 44. Flaherty AW, Graybiel AM. Input-output organization of the sensorimotor striatum in the squirrel monkey. J Neurosci. 1994;14:599–610. 45. Aldridge JW, Anderson RJ, Murphy JT. Sensory-motor processing in the caudate nucleus and globus pallidus: a single-unit study in behaving primates. Can J Physiol Pharmacol. 1980;58:1192–1201. 46. Kimura M. The role of primate putamen neurons in the association of sensory stimulus with movement. Neurosci Res. 1986;3:436–443. 47. McFarland NR, Haber SN. Convergent inputs from thalamic motor nuclei and frontal cortical areas to the dorsal striatum in the primate. J Neurosci. 2000;20:3798–3813. 48. Hedreen JC, DeLong MR. Organization of striatopallidal, striatonigal, and nigrostriatal projections in the macaque. J Comp Neurol. 1991;304:569–595. 49. Haber SN, Lynd E, Klein C, Groenewegen HJ. Topographic organization of the ventral striatal efferent projections in the rhesus monkey: an anterograde tracing study. J Comp Neurol. 1990;293:282–298. 50. Selemon LD, Goldman-Rakic PS. Topographic intermingling of striatonigral and striatopallidal neurons in the rhesus monkey. J Comp Neurol. 1990;297:359–376. 51. Lynd-Balta E, Haber SN. Primate striatonigral projections: a comparison of the sensorimotor-related striatum and the ventral striatum. J Comp Neurol. 1994;345:562–578. 52. Strick PL. Anatomical analysis of ventrolateral thalamic input to primate motor cortex. J Neurophysiol. 1976;39:1020–1031. 53. Ilinsky IA, Jouandet ML, Goldman-Rakic PS. Organization of the nigrothalamocortical system in the rhesus monkey. J Comp Neurol. 1985;236:315–330. 54. Kuo J, Carpenter MB. Organization of pallidothalamic projections in the rhesus monkey. J Comp Neurol. 1973;151:201–236. 55. McFarland NR, Haber SN. Thalamic connections with cortex from the basal ganglia relay nuclei provide a mechanism for integration across multiple cortical areas. J Neurosci. 2002; 22: 8117–8132. 56. Mink JW. The basal ganglia: focused selection and inhibition of competing motor programs. Prog Neurobiol. 1996;50: 381–425. 57. Wise SP, Murray EA, Gerfen CR. The frontal cortex-basal ganglia system in primates. Crit Rev Neurobiol. 1996;10:317–356.

48

NEUROANATOMY

58. Cools R, Clark L, Robbins TW. Differential responses in human striatum and prefrontal cortex to changes in object and rule relevance. J Neurosci. 2004;24:1129–1135. 59. Muhammad R, Wallis JD, Miller EK. A comparison of abstract rules in the prefrontal cortex, premotor cortex, inferior temporal cortex, and striatum. J Cogn Neurosci. 2006;18:974–989. 60. Hikosaka O, Miyashita K, Miyachi S, Sakai K, Lu X. Differential roles of the frontal cortex, basal ganglia, and cerebellum in visuomotor sequence learning. Neurobiol Learning Memory. 1998;70:137–149. 61. Calzavara R, Mailly P, Haber SN. Relationship between the corticostriatal terminals from areas 9 and 46, and those from area 8A, dorsal and rostral premotor cortex and area 24c: an anatomical substrate for cognition to action. Eur J Neurosci. 2007;26:2005–2024. 62. Kasanetz F, Riquelme LA, Della-Maggiore V, O’Donnell P, Murer MG. Functional integration across a gradient of corticostriatal channels controls UP state transitions in the dorsal striatum. Proc Natl Acad Sci USA. 2008;105:8124–8129. 63. Satoh T, Nakai S, Sato T, Kimura M. Correlated coding of motivation and outcome of decision by dopamine neurons. J Neurosci. 2003;23:9913–9923. 64. Pagnoni G, Zink CF, Montague PR, Berns GS. Activity in human ventral striatum locked to errors of reward prediction. Nat Neurosci. 2002;5:97–98. 65. Redgrave P, Gurney K. The short-latency dopamine signal: a role in discovering novel actions? Nat Rev Neurosci. 2006;7:967–975. 66. Volkow ND, Wang GJ, Telang F, Fowler JS, Logan J, Childress AR, Jayne M, Ma Y, Wong C. Cocaine cues and dopamine in dorsal striatum: mechanism of craving in cocaine addiction. J Neurosci. 2006;26:6583–6588. 67. Porrino LJ, Smith HR, Nader MA, Beveridge TJ. The effects of cocaine: a shifting target over the course of addiction. Prog Neuropsychopharmacol Biol Psychiatry. 2007;31: 1593–1600. 68. Everitt BJ, Robbins TW. Neural systems of reinforcement for drug addiction: from actions to habits to compulsion. Nat Neurosci. 2005;8:1481–1489. 69. Hokfelt T, Martensson R, Bjorklund A, Kleinau S, Goldstein M. Distributional maps of tyrosine-hydroxylase immunoreactive neurons in the rat brain. In: Bjorklund A, Hokfelt T, eds. Handbook of Chemical Neuroanatomy, Vol. II: Classical Neurotransmitters in the CNS, Part I. Amsterdam: Elsevier; 1984:277–379. 70. Olszewski J, Baxter D. Cytoarchitecture of the Human Brain Stem. 2nd ed. Basel: S. Karger; 1982. 71. Haber SN, Ryoo H, Cox C, Lu W. Subsets of midbrain dopaminergic neurons in monkeys are distinguished by different levels of

72.

73.

74.

75.

76.

77.

78.

79.

80.

81.

82.

83.

84.

mRNA for the dopamine transporter: Comparison with the mRNA for the D2 receptor, tyrosine hydroxylase and calbindin immunoreactivity. J Comp Neurol. 1995;362:400–410. Lavoie B, Parent A. Dopaminergic neurons expressing calbindin in normal and parkinsonian monkeys. Neuroreport. 1991;2(10):601–604. Haber SN, Gdowski MJ. The basal ganglia. In: Paxinos G, Mai JK, eds. The Human Nervous System. 2nd ed. New York, NY: Elsevier Press; 2004:677–738. Fudge JL, Haber SN. Bed nucleus of the stria terminalis and extended amygdala inputs to dopamine subpopulations in primates. Neuroscience. 2001;104:807–827. Lynd-Balta E, Haber SN. The organization of midbrain projections to the striatum in the primate: sensorimotor-related striatum versus ventral striatum. Neuroscience. 1994;59: 625–640. Parent A, Mackey A, De Bellefeuille L. The subcortical afferents to caudate nucleus and putamen in primate: a fluorescence retrograde double labeling study. Neuroscience. 1983;10(4): 1137–1150. Watanabe K, Lauwereyns J, Hikosaka O. Neural correlates of rewarded and unrewarded eye movements in the primate caudate nucleus. J Neurosci. 2003;23:10052–10057. Takikawa Y, Kawagoe R, Hikosaka O. Reward-dependent spatial selectivity of anticipatory activity in monkey caudate neurons. J Neurophysiol. 2002; 87:508–515. Tanaka SC, Doya K, Okada G, Ueda K, Okamoto Y, Yamawaki S. Prediction of immediate and future rewards differentially recruits cortico-basal ganglia loops. Nat Neurosci. 2004;7:887–893. Zheng T, Wilson CJ. Corticostriatal combinatorics: the implications of corticostriatal axonal arborizations. J Neurophysiol. 2002;87:1007–1017. Mallet N, Le Moine C, Charpier S, Gonon F. Feedforward inhibition of projection neurons by fast-spiking GABA interneurons in the rat striatum in vivo. J Neurosci. 2005;25: 3857–3869. Porrino LJ, Lyons D, Smith HR, Daunais JB, Nader MA. Cocaine self-administration produces a progressive involvement of limbic, association, and sensorimotor striatal domains. J Neurosci. 2004; 24:3554–3562. Lehericy S, Benali H, Van de Moortele PF, Pelegrini-Issac M, Waechter T, Ugurbil K, Doyon J. Distinct basal ganglia territories are engaged in early and advanced motor sequence learning. Proc Natl Acad Sci USA. 2005;102:12566–12571. Belin D, Everitt BJ. Cocaine seeking habits depend upon dopamine-dependent serial connectivity linking the ventral with the dorsal striatum. Neuron. 2008;57:432–441.

2.4

TheRelationship betweenDopaminergicAxons andGlutamatergic Synapses in theStriatum:StructuralConsiderations JONATHAN MOSS AN D J. PAUL B OL A M

INTRODUCTION The basal ganglia are a group of highly interconnected nuclei involved in a variety of functions including movement and cognition. The dorsal component, which is primarily related to motor and associative functions, consists of the striatum, external segment of the globus pallidus, subthalamic nucleus, internal segment of the globus pallidus, and substantia nigra pars reticulata. The last two structures form the output nuclei of the basal ganglia (Fig, 2.4.1). The major inputs to the basal ganglia arise in the cerebral cortex and thalamus and are carried by the corticostriatal and thalamostriatal pathways. This information is processed in the striatum and transmitted by various routes to the output nuclei. The basal ganglia then influence behavior by these structures projecting to thalamus and thence to the cortex or to other subcortical structures involved in movement. Overlying this feedforward system of the basal ganglia is feedback from dopamine neurons in the substantia nigra pars compacta (SNc). These neurons massively innervate the striatum and also provide innervation of other regions of the basal ganglia, albeit at a much lower density. At the level of the striatum, the principal role of the dopaminergic innervation is to modulate the flow of cortical and thalamic information through the basal ganglia. The objective of this brief review is to summarize data relating principally to the anatomical substrate of the interaction between both glutamatergic corticostriatal synapses and thalamostriatal synapses with dopaminergic axons and terminals in the striatum.

GENERAL ASPECTS OF DOPAMINERGIC INNERVATION OF THE STRIATUM Dopamine neurons of the SNc account for a remarkably small number of neurons. It is estimated that there are 7000–8000 neurons1,2 in each SNc of the rat, with a total of 12,000 in the entire SN and about another

20,000 dopamine neurons in the ventral tegmental area (VTA).2 The remarkable nature of these neurons lies not only in their numbers but also in their innervation of the forebrain. It is well known that the density and distribution of markers of dopaminergic neurons and transmission in the striatum are the highest in the brain,3,4 but although many studies have examined the somatodendritic properties and locations of individual dopamine neurons filled in vivo, it was not until very recently that the axonal field of individual dopamine neurons was revealed.5 Analysis of individual dopamine neurons (revealed by infection with a viral vector expressing membrane-targeted green fluorescent protein) in the rat brain showed that on average the total length of the axon in the striatum is in the region of 47 cm and the arborization can extend to occupy up to 5.7% of the volume of the striatum.5 The remarkable length of the axons of individual neurons is reflected in the estimates of the number of dopaminergic synapses formed in the striatum. Based on the number of neurons in the SNc and the striatum1,2 and the known synaptic organization of the dopaminergic nigrostriatal projection,6 we estimate that an individual dopaminergic neuron gives rise to between 170,00 and 408,000 synapses in the striatum (Table 2.4.1). This figure is close to the estimate of Wickens and Arbuthnott,7 who, using a completely different approach and set of assumptions based on densities of synapses and neurons, concluded that individual dopamine neurons give rise to about 370,000 synapses in the striatum.7 Furthermore, these figures are close to the estimates of Anden et al.3 of a total axon length of 30 cm and 250,000 varicosities per SNc dopaminergic neuron, based on the analysis of tissue stained by the histofluorescence method3 (cited by, and figures recalculated by, Bjo¨rklund and Lindvall4). To put these figures in perspective, data from rats have indicated that neurons of the external globus pallidus give rise to approximately 2000 synapses,8,9 striatal spiny neurons probably give rise 49

50

NEUROANATOMY

Cortex

Striatum

STN

Thalamus

GPe

SNc

Brainstem Spinal Cord

SNr/GPi

Output or feedback

Simplified block diagram of the basal ganglia showing the principal connections of dopamine neurons. The nuclei of the basal ganglia (included in the light gray box) consist of the striatum, the external segment of the globus pallidus (GPe), the subthalamic nucleus (STN), the substantia nigra pars reticulata and the internal segment of the globus pallidus (SNr/GPi), and the substantia nigra pars compacta (SNc). The two major inputs to the basal ganglia are from the cortex and the thalamus (mainly the intralaminar nuclei). The SNr and GPi constitute the output nuclei of the basal ganglia projecting to the thalamus and thence back to the cortex or to other subcortical structures. Dopamine neurons of the SNc provide massive feedback innervation of the striatum but also of other regions of the basal ganglia plus the prefrontal cortex. Dopamine neurons may also modulate neurons of the SNr by the dendritic release of dopamine.

FIGURE 2.4.1.

to about 300 synapses,10,11 and fast-spiking interneurons in the striatum give rise to about 5000 synapses.12,13 It should be noted that the figures for dopamine neurons are also probably underestimates, as individual dopamine neurons innervate multiple regions of the basal ganglia, where they form synapses and release dopamine (see, for instance, 5,14,15). Whatever the precise figures for dopamine neurons are, they are remarkable neurons when compared to classical central nervous system (CNS) neurons. Such a large axonal arborization raises questions about the control of the activity of individual boutons, which may be as far as several tens of millimeters from the site of initiation of the axon potential at the axon hillock or proximal dendrite. Do all axon potentials invade all of the extensive and tortuous branches of the axonal arbor? Does such a large axonal arbor, which requires supply and support, render the neuron particularly susceptible to stressors that lead to cell death? It

is interesting to note that we estimate, using similar methods and assumptions as described in Table 1, that dopaminergic neurons in the VTA, which are less susceptible to dying in Parkinson’s disease, give rise to far less synapses (in the range of 12,000–30,000 per neuron).

SYNAPTIC ORGANIZATION OF THE DOPAMINE INNERVATION OF THE STRIATUM At the level of the striatum, axons of dopaminergic neurons give rise to small vesicle-containing varicosities that form small, mainly symmetrical (Gray’s Type 2) synapses (Fig. 2.4.2). The synapses are often difficult to visualize because of the small size of the specialization and the fact that their integrity is easily lost in suboptimally fixed tissue. Nonsynaptic segments of dopaminergic axons may also contain vesicles, and synapses may be formed by nonvaricose segments of the axons.16,17 Most studies in the rat agree that one of the principal synaptic targets of dopaminergic terminals in the striatum are dendritic spines (51%–65% of synapses formed by dopaminergic terminals are with spines).16,18–20 The remaining synapses are with dendritic shafts (30%–46%) and perikarya (2%– 6%). One study in the rat, however, identified a smaller proportion in contact with spines (30%) and a correspondingly higher proportion in contact with dendritic shafts (67%).21 Interestingly, there is very little difference in the distribution of synaptic targets between the patch/striosome and matrix compartments of the striatum.19 In primates (squirrel monkey), it appears that there is greater heterogeneity in the type of synaptic specialization (i.e., a greater proportion form asymmetrical synapses), and only 22.5% of synapses were identified to be in contact with spines and 72% in contact with dendritic shafts.22 It is commonly the case that the synaptic contacts with dendritic spines are associated with the necks of spines that are also in synaptic contact with a terminal forming an asymmetrical, presumably excitatory, synaptic contact (Gray’s Type 1),16,22 and indeed, some have been shown to be derived from the cortex22,23 (see below and Fig. 2.4.3A).

SYNAPTIC ORGANIZATION OF THE CORTICAL INNERVATION OF THE STRIATUM The corticostriatal projection is both bilateral and topographical in nature, and several organizational principles have been described that contribute to a complex

2.4: THE RELATIONSHIP BETWEEN DOPAMINERGIC AXONS AND GLUTAMATERGIC SYNAPSES TABLE 2.4.1.

51

Number of Synapses Formed by a Single Dopamine Neuron in the Striatum

1. Average number of dendritic spines on one MSN

6250–15,000

2. Percentage of axo-spinous synapses that involve cortical and thalamic terminals

64.8%

3. Average number of dendritic spines postsynaptic to cortical or thalamic terminals on one MSN (1) x (2)

4050–9720

4. Percentage of these dendritic spines in synaptic contact with a dopamine terminal

6.666%

5. Number of these dendritic spines on one MSN in synaptic contact with a dopamine terminal (3) x (4)

270–648

6. Percentage of dopamine terminals that contact dendritic spines (not shafts or cell bodies)

61.3%

7. Multiplying factor to incorporate synapses with dendritic shafts and cell bodies; reciprocal of 0.613 (6)

1.632

8. Number of dopamine terminals forming synapses with one MSN (5) x (7)

441–1058

9. Number of MSNs in the striatum (one hemisphere)

2,780,000

10. Number of dopamine terminals forming synapses with all MSNs in one hemisphere (8) x (9)

1,224,979,200–2,939,950,080

11. Number of dopamine neurons in the substantia nigra pars compacta (one hemisphere)

7200

12. Number of symmetrical synapses formed by one dopamine neuron in the rat striatum (10)/(11)

170,136–408,326

Note that to arrive at the number of dopaminergic synapses in contact with an individual MSN, we have used quantitative data from spines that are postsynaptic to cortical or thalamic terminals. This will introduce error because dopaminergic terminals may contact other spines and for the reason indicated below. Value in 1 is the range indicated by Kincaid et al.45 Value in 2 is from Lacey et al.37 This figure represents the percentage of axo-spinous synapses involving VGluT1-positive (cortical) and VGluT2-positive (thalamic) terminals. It is likely to be an underestimate due to false-negative labeling of terminals. Values in 4 and 6 are from Moss and Bolam.6 Values in 9 and 11 are from Oorschot.1 The figure of 7200 dopamine neurons may be an underestimate of the true number of nigro-striatal dopaminergic neurons, as dopamine neurons located in the pars reticulata also project to the striatum.

and heterogeneous projection. On the basis of single cell filling, several classes of corticostriatal neurons have been described that differ in their cortico-cortical and cortico-fugal projections as well as in their pattern of innervation of the striatum (see 24). The corticostriatal projection is also heterogeneous with respect to the patch/striosome and matrix subdivisions of the striatum.24–32 Limbic cortical areas show selectivity for the innervation of the patch/striosome, whereas other areas show selectivity for the matrix. Furthermore, neurons in deep layer 5 and layer 6 of the cortex selectively innervate the patch/striosome, whereas neurons located in upper layer 5 and layer 3 selectively innervate the matrix. The modular organization of the corticostriatal termination within the matrix, referred to as matrisomes, represents a further level of heterogeneity.27,32–35

Terminals in the striatum that are derived from the cortex form asymmetrical (Gray’s Type 1) synaptic specializations (Fig. 2.4.3A; for references see 36). Ultrastructural analysis of corticostriatal terminals labeled by anterograde degeneration, anterograde tracing, or immunolabeling for vesicular glutamate transporter type 1 (VGluT1; see below)36–39 reveals that a high proportion (>95%) make synaptic contact with dendritic spines. Since medium-sized spiny projection neurons (MSNs) account for the majority of spines in the striatum, this observation suggests that they are likely to be the major targets of the cortical projection, and this is supported by direct analysis of MSNs.40–42 Furthermore, spiny neurons giving rise to both the direct and indirect pathways of information flow through the basal ganglia receive input from the cortex.43 Corticostriatal terminals originating from the

FIGURE 2.4.2. Dopaminergic neurons give rise to small, symmetrical synapses in the striatum. TH-positive terminals (TH) making symmetrical synaptic contact (arrows) with a dendritic shaft in (A) and dendritic spines in (B–D). Scale bars: 200 nm.

52

NEUROANATOMY

FIGURE 2.4.3. Synaptic targets of cortical and thalamic terminals in the striatum. (A) An axon terminal in the monkey putamen, anterogradely labeled from the cortex (Ctx), makes asymmetrical synaptic contact (arrowhead) with a dendritic spine (s). A THpositive axon terminal (TH) makes symmetrical synaptic contact (small arrow) with the same spine. Dendritic spines are the main synaptic target of corticostriatal terminals. (B) An axon terminal in the rat striatum derived from a neuron in the central lateral nucleus of the thalamus (CL) makes asymmetrical synaptic contact (arrowhead) with a dendritic spine (s) that can be seen to arise from a dendritic shaft (d). Most terminals derived from the CL make asymmetrical axospinous synapses. (C) An axon terminal in the rat striatum derived from a neuron in the parafascicular nucleus of the thalamus (Pf) makes asymmetrical synaptic contact with a dendritic shaft (d). About 63%–89% of terminals from the parafascicular nucleus make synaptic contact with dendritic shafts and the remainder with dendritic spines. The axon terminals in (A) and (C) were derived from neurons that were recorded and juxtacellularly labeled in vivo. Scale bars: 200 nm. Source: Data in (A) derived and modified from Smith et al.22 Data in (B) and (C) derived and modified from Lacey et al.65

contralateral cortex form contact more frequently with dendritic spines,43 and the two broad classes of corticostriatal neurons,24 that is, those that project preferentially within the telencephalon and corticopyramidal neurons that give rise to a collateral to the striatum, have different patterns of innervation of the striatum.44 The former give rise to relatively small axonal boutons that preferentially innervate the spines of spiny neurons giving rise to the direct pathway, whereas the latter give rise to relatively large boutons that preferentially innervate the spines of spiny neurons giving rise to the indirect pathway. Quantitative analysis of the pattern of innervation of the striatum by individual cortical axons suggests that individual spiny neurons are likely to receive only about four synapses from an individual corticostriatal axon45,46; there is thus a high degree of divergence of cortical axons in the striatum and a high degree of convergence at the single striatal cell level. Striatal interneurons that express the calcium-binding protein, parvalbumin, often referred to as fast-spiking interneurons,13,47–50 are prominently innervated by cortical axons,51,52 but the pattern of innervation is different from that of cortical input to MSNs. Individual corticostriatal axons (which may arise from functionally diverse regions of the cortex) make multiple synaptic contacts

with individual parvalbumin-positive GABAergic interneurons.53 The population of striatal GABAergic interneurons that express somatostatin and neuropeptide Y immunoreactivity and nitric oxide synthase have also been shown to receive synaptic input to their dendrites from corticostriatal terminals.54 Although electrophysiological analyses indicate that the cholinergic neurons readily respond to cortical stimulation,55,56 analysis of choline acetyltransferase (ChAT)–immunostained structures in striatal tissue containing terminals anterogradely labeled from frontal cortex has failed to identify a cortical input.52 This situation is similar in the nucleus accumbens with respect to the hippocampal input.57 These findings indicate that cholinergic neurons receive little, if any, synaptic input from the frontal cortex in their proximal regions, that is, the regions of the neurons that were labeled by ChAT immunocytochemistry. It is, of course, possible that other cortical regions make synaptic contact with spiny neurons or indeed that the cortical input occurs in the most distal regions of the dendritic tree that were not immunostained (but see reference Thomas et al.57a).

SYNAPTIC ORGANIZATION OF THALAMIC INNERVATION OF THE STRIATUM The thalamostriatal projection originates mainly from the intralaminar thalamic nuclei, although minor projections arise in other thalamic nuclei (see 58,59). Similar to corticostriatal projections, the thalamostriatal projections are topographically organized and show heterogeneities in relation to the morphology of the projecting axons, their distribution with respect to the patch/striosome-matrix organization of the striatum, and the distribution of postsynaptic targets in the patches/striosomes and matrix.38,60–63 Singlecell labeling and tracing studies have identified at least two axonal morphologies of the thalamostriatal projection.61,64–66 Ultrastructural analyses of the thalamostriatal system have shown that terminals derived from the thalamus are similar in morphology to cortical terminals, they form asymmetrical synaptic specializations, they are packed with vesicles and they usually contain one or two mitochondria (Fig. 2.4.3B,C).22,38,57,62,65–71 When considering the projection as a whole, by the analysis of terminals immunolabeled for VGluT2 (see below), it is apparent that the principal targets of the thalamostriatal projections, like those of the corticostriatal projections, are dendritic spines of MSNs. Approximately 60%–75% of VGluT2-positive terminals making synaptic contact with the spines and about 25%–40% making contact

2.4: THE RELATIONSHIP BETWEEN DOPAMINERGIC AXONS AND GLUTAMATERGIC SYNAPSES

with dendritic shafts have been reported.6,37–39 However, analysis of projections from individual different subnulcei of the thalamus by anterograde labeling38,66 or by juxtacellular labeling of individual thalamostriatal neurons65 has revealed that the proportion of terminals contacting spines or dendrites is related to the nucleus from which the projection originates. Hence, boutons derived from the parafascicular nucleus terminate primarily on dendritic shafts,65,66,71,72 but the precise ratio of spines to shafts varies among individual neurons.65 Other nuclei giving rise to thalamostriatal projections principally target dendritic spines,38,65,66,68 and it has been proposed that neurons in the centromedian nucleus of the thalamus, at least, preferentially target MSNs that give rise to the direct pathway.73 Variability also exists in the ratio of dendritic spines to shafts when considering the patch/striosomes and matrix subcompartments of the striatum.38,39,62 In addition to MSNs, thalamostriatal neurons innervate interneurons. Cholinergic interneurons, which possess large perikarya and long, essentially spine-free dendrites, receive asymmetrical synaptic input from terminals derived from the parafascicular nucleus of the thalamus.67,74 A similar arrangement exists in the nucleus accumbens.57 Parvalbumin-expressing and neuropeptide Y–expressing, but not calretinin-expressing, GABA interneurons have been shown to receive thalamic input in rat and monkey74,75 (but see 76).

INTERACTIONS BETWEEN DOPAMINE AND GLUTAMATE IN STRIATUM Central to our understanding of basal ganglia function is the concept that the role of dopamine is to modulate transmission at glutamatergic synapses within the striatum. This has traditionally been considered to be a modulatory effect on corticostriatal synapses, but it is also likely to be an effect on thalamostriatal synapses (see below).77–80 This modulatory effect, or interaction between dopaminergic transmission and glutamatergic transmission, takes many forms. Pharmacological analyses have shown that dopamine can directly influence the release of glutamate at glutamatergic synapses within the striatum. Plasticity of corticostriatal synapses in the form of long-term potentiation and long-term depression is dependent on many factors, including dopamine acting upon the D1 or D2 subtypes of dopamine receptors.78–85 Plasticity of thalamostriatal synapses may also be dependent on released dopamine.72,86,87 Furthermore, the loss of dopamine innervation of the striatum leads not only to a loss of spines but also to a loss of excitatory synapses.88–90

53

The anatomical substrates of such interactions have long been considered to be the convergent input of excitatory synapses at the head of dendritic spines of MSNs and dopaminergic synapses at the neck of the spines (see Fig. 2.4.3A).16,22,23,68,91 This synaptic relationship has been identified for corticostriatal synapses in the dorsal striatum22,23 and thalamostriatal synapses (derived from the paraventricular nucleus of the thalamus) in the ventral striatum,92 although evidence from tract tracing studies is lacking in the dorsal striatum22 (but see below). This triadic arrangement of excitatory input at the head of the dendritic spine and dopamine input at the neck has also been identified in other regions of the brain, including cortex93 and amygdala.94 Thus, dopamine acting upon receptors at the necks of spines influences the intracellular signaling pathways initiated by the activation of glutamate receptors at the heads of spines, thereby modulating the responsiveness of the postsynaptic spine to the released glutamate.

QUANTITATIVE ANALYSIS OF THE DOPAMINERGIC INNERVATION OF THE STRIATUM REVEALS THE PRINCIPLES OF INNERVATION Quantitative analysis of possible sites of interaction between dopaminergic and glutamatergic synapses based on anterograde labeling studies or combined anterograde labeling and immunocytochemical studies is limited by the problems of false-negative labeling of terminals. Anterograde tracing will label only a small proportion of the population of terminals in a pathway. These problems, in relation to the quantitative analysis of the corticostriatal and thalamostriatal pathways, have been overcome by the discovery (that Na+-dependent inorganic phosphate transporters that act as VGluTs)95–98 selectively label corticostriatal and thalamostriatal terminals, respectively37–39,72,99 (but see 100,101). Immunocytochemical analyses using antibodies against these transporters have enabled large parts, if not the whole, of these projections to be studied, and have found that the percentage of axon terminals in the striatum that are derived from the thalamus (25% of asymmetrical synapses) is of the same order of magnitude as that from the cortex (35% of asymmetrical synapses).37 We have taken advantage of these markers to define quantitatively the spatial relationship between corticostriatal terminals and dopaminergic axons, as well as that between thalamostriatal terminals and dopaminergic axons.6 Quantitative electron microscopic analysis was performed on sections of rat striatum immunolabeled to

54

NEUROANATOMY

reveal tyrosine hydroxylase, as a marker dopaminergic axons, and either VGluT1 or VGluT2 as markers of corticostriatal and thalamostriatal terminals, respectively (Fig. 2.4.4). The essential findings of the study were as follows:

• The majority of cortical terminals made synaptic

contact with dendritic spines (96%), and 20% of the postsynaptic spines were apposed by a dopaminergic axon. In 9% of the cases, the postsynaptic structure received synaptic input from the dopaminergic axon. • Like the cortical terminals, the majority of thalamic terminals made synaptic contact with dendritic spines (71%) and, similar to the cortical terminals, 27% of the structures postsynaptic to the thalamic terminals (spines and dendrites) were also apposed by a dopaminergic axon. In 9% of the cases, the dopaminergic axon formed a synapse with the structure postsynaptic to the thalamic terminal. • Randomly selected cellular profiles within the striatum, when corrected for the length of their perimeter within the electron micrographs, have a probability of being apposed by, or in synaptic contact with, a dopaminergic axon similar to that of the spines and dendrites postsynaptic to cortical and thalamic terminals. • Similarly, glutamatergic synaptic terminals from the cortex or thalamus, when corrected for their size, have a probability of being apposed by a dopaminergic axon similar to that of the structures postsynaptic to them. These results demonstrate that a proportion of those spines and dendrites postsynaptic to cortical terminals receive synaptic input from dopaminergic terminals, and that this is also the case for structures postsynaptic to thalamic terminals (Fig. 2.4.5). There are several important implications of these observations:

• The anatomical substrate for the interaction of

dopamine and glutamate applies equally to corticostriatal and thalamostriatal synapses. • The plasticity of thalamostriatal synapses is therefore likely to have the same degree of dependency upon dopamine as the plasticity of corticostriatal synapses. • Since the frequency of the spatial relationship between an excitatory synapse and a dopaminergic synapse was the same for randomly selected cellular profiles and a dopaminergic synapse, the relationship between dopaminergic synapses and glutamatergic synapses is unlikely to be a selective

The spatial relationship between excitatory synapses and dopaminergic synapses in the striatum. Corticostriatal terminals were revealed by immunogold labeling for VGluT1, thalamostriatal terminals by immunogold labeling for VGluT2, and dopaminergic axons by immunoperoxidase labeling for tyrosine hydroxylase (TH). (A) A VGluT1-positive bouton (b; corticostriatal) makes asymmetrical synaptic contact (arrowhead) with the head of a long, thin spine (s). A TH-positive terminal (TH; dopaminergic) makes symmetrical synaptic contact (arrow) with the neck of the same spine. (B) A VGluT1-positive bouton (b; corticostriatal) makes asymmetrical synaptic contact with a dendritic shaft (d) that is apposed (arrow) by a TH-positive axon (TH; dopaminergic). Note the additional corticostriatal terminal forming a synapse with a spine at the bottom right of this micrograph. (C) A VGluT2-positive bouton (b; thalamostriatal) makes asymmetrical synaptic contact (arrowhead) with a spine (s) that arises from a dendritic shaft (d). A TH-positive bouton (TH; dopaminergic) makes symmetrical synaptic contact (arrows) with both the spine (s) and the dendritic shaft (d). (D) A VGluT2-positive bouton (b; thalamostriatal) makes asymmetrical synaptic contact (arrowhead) with a dendritic shaft (d) that is in symmetrical synaptic contact (arrow) with a THpositive terminal (TH; dopaminergic). Scale bars: 200 nm. Source: Data modified from Moss and Bolam.6

FIGURE 2.4.4.

or targeted phenomenon. The chance of a striatal structure being apposed by, or in synaptic contact with, a dopaminergic axon seems to be solely dependent on the size of the structure. The spatial relationship between dopaminergic axons and glutamatergic synapses simply relates to the size of the presynaptic and postsynaptic structures, not to their phenotype. • If the role of dopamine in the modulation of glutamatergic transmission is so critical to our understanding of the function of the striatum and the basal ganglia in general, then the question arises as to why such a small proportion (9%) of the

2.4: THE RELATIONSHIP BETWEEN DOPAMINERGIC AXONS AND GLUTAMATERGIC SYNAPSES

structures postsynaptic to glutamatergic synapses also receive synaptic input from a dopaminergic terminal. Are the other glutamatergic synapses in the striatum not modulated by dopamine? One possible explanation is that, in addition to synaptic transmission, dopaminergic transmission may also occur by volume transmission as a consequence of spillover of synaptically released dopamine or the release of dopamine at nonsynaptic sites.102–106 The ‘‘sphere of influence’’ of released dopamine is likely to depend on many factors, including quantal size and the density and distribution of dopamine transporters and receptors. It has been proposed that the sphere of influence of dopamine spillover in a concentration sufficient to stimulate dopamine receptors has a radius of 2–8 mm.103 A precise synaptic relationship between a glutamatergic terminal and a dopaminergic terminal may thus not be necessary for released dopamine to modulate the strength of a glutamatergic synapse. In order to address this and to see, on average, how close dopaminergic structures are to glutamatergic synapses, we examined the proximity of dopaminergic axons and synapses to glutamatergic synapses.6 Using the same data set described above, we found that every glutamatergic synapse is within 0.5 mm of a dopaminergic axon and within about 1 mm of a dopaminergic synapse (Table 2.4.2). In view of the estimates of the distance that dopamine may diffuse from the synapse, these findings suggest that every structure, including glutamatergic synapses, will be within overlapping spheres of influence of synaptically released dopamine. Thus, all glutamatergic synapses are likely to be within reach of a concentration of dopamine high enough to stimulate both high- and low-affinity receptors.103 Efficacy of transmission will thus depend on the density and distribution of extrasynaptic dopamine receptors43,107–109 and, of course, will have different temporal characteristics to synaptic transmission. It should be noted that, as one would predict from the analysis described above, all randomly selected cellular profiles within the striatum are also located within about 0.5 mm of a dopaminergic axon and within about 1 mm of a dopaminergic synapse (Table 2.4.2). This reinforces the idea that there is no selectivity in the dopaminergic nigrostriatal pathway; rather, the potential for functional connectivity is dependent on the size of the target structure, the density of the projection, and the particular axon involved.5 Functional connectivity thus depends almost entirely on the density, distribution, and location of dopamine receptors.

55

Cortex 96% 9%

MSN

10%

Random Structures

9% 71%

Thalamus

FIGURE 2.4.5. Summary diagram of the convergence of glutamatergic and dopaminergic signals in the striatum and its nonselective nature. Cortical and thalamic afferents to the striatum (red) make asymmetrical synaptic contact with dendritic structures (blue) of a medium-sized spiny projection neuron (MSN, white). The majority of these contacts are with dendritic spines (cortical, 96%; thalamic, 71%), of which 9% receive a second input from a dopaminergic axon from the substantia nigra pars compacta (yellow). This, however, is no different from the proportions of random striatal structures (green) contacted by dopaminergic axons (10%), which demonstrates the nonselective nature of the relationship. In addition, dopamine (yellow clouds) spill over from the synapse and diffuse in concentrations capable of activating dopamine receptors for up to 8 mm. (See Color Plate 2.4.5.)

CONCLUDING COMMENTS Dopamine neurons are remarkable in their complexity: a small population of neurons gives rise to a phenomenally dense innervation of the striatum, and individual neurons have vast axonal arbors that give rise to hundreds of thousands of synapses. The organization of what is central to basal ganglia function (i.e., the interaction between dopamine and glutamate) is such that striatal neurons are embedded in a dense network of dopamine axons and every structure has a similar probability of being apposed by, or in synaptic contact with, a dopaminergic axon. Furthermore, every structure in the striatum is within overlapping spheres of

56

NEUROANATOMY

TABLE 2.4.2. The Proximity of Cortical Synapses, Thalamic Synapses, and Random Points in the Striatum to Dopaminergic Axons and Synapses

Proximity to Dopaminergic Axons

Proximity to Dopaminergic Synapses

Proportion within 0.5 m

Average distance between

Proportion within 0.5 m

Average distance between

Cortical Synapses

108%

0.49 mm

20%

0.85 mm

Thalamic Synapses

104%

0.49 mm

10%

1.08 mm

Random Structures

96%

0.51 mm

11%

1.04 mm

Source: Data derived from the quantitative analysis of Moss and Bolam.6

influence of synaptically released dopamine that may spill over and diffuse from the synapse. These structural characteristics thus underlie the phasic actions of dopamine at synapses, presumably in response to bursts of activity of dopamine neurons. They also underlie the tonic effects of dopamine, which are likely to occur as a consequence of tonic release at synapses, as well as the diffuse spillover of dopamine from synapses and possibly nonsynaptic sites. Given these structural characteristics, the critical factor in the expression of tonic dopamine function is not the specific location of dopamine synapses, but rather the distribution and density of dopamine receptors that, like most metabotropic receptors, are located at both synaptic and extrasynaptic sites.43,107–109

ACKNOWLEDGMENTS The authors’ work described in this chapter was supported by the Medical Research Council, The European community (FP7 project number 201716) and The Parkinson’s Disease Society (UK). JM was in receipt of a Medical Research Council studentship.

REFERENCES 1. Oorschot DE. Total number of neurons in the neostriatal, pallidal, subthalamic, and substantia nigral nuclei of the rat basal ganglia: a stereological study using the Cavalieri and optical disector methods. J Comp Neurol.1996;366(4):580– 599. 2. Nair-Roberts RG, Chatelain-Badie SD, Benson E, White-Cooper H, Bolam JP, Ungless MA. Stereological estimates of dopaminergic, GABAergic and glutamatergic neurons in the ventral tegmental area, substantia nigra and retrorubral field in the rat. Neuroscience. 2008;152(4):1024–1031. 3. Anden NE, Fuxe K, Hamberger B, Ho¨kfelt T. A quantitative study on the nigro-neostriatal dopamine neuron system in the rat. Acta Physiol Scand. 1966;67(3):306–312.

4. Bjo¨rklund A, Lindvall O. Dopamine-containing systems in the CNS. In: Bjo¨ rklund A, Ho¨ kfelt T, eds. Handbook of Chemical Neuroanatomy. Amsterdam: Elsevier; 1984:2: 55–122. 5. Matsuda W, Furuta T, Nakamura KC, et al. Single nigrostriatal dopaminergic neurons form widely spread and highly dense axonal arborizations in the neostriatum. J Neurosci. 2009;29(2):444–453. 6. Moss J, Bolam JP. A dopaminergic axon lattice in the striatum and its relationship with cortical and thalamic terminals. J Neurosci. 2008;28(44):11221–11230. 7. Wickens J, Arbuthnott GW. Structural and functional interactions in the striatum at the receptor level. In: Dunnett SB, Bentivoglio M, Bjo¨rklund A, Ho¨kfelt T, eds. Handbook of Chemical Neuroanatomy. Amsterdam: Elsevier; 2004:21: 199–236. 8. Bevan MD, Clarke NP, Bolam JP. Synaptic integration of functionally diverse pallidal information in the entopeduncular nucleus and subthalamic nucleus in the rat. J Neurosci. 1997;17(1):308–324. 9. Kita H, Kitai ST. The morphology of globus pallidus projection neurons in the rat: an intracellular staining study. Brain Res. 1994;636:308–319. 10. Kawaguchi Y, Wilson CJ, Emson PC. Projection subtypes of rat neostriatal matrix cells revealed by intracellular injection of biocytin. J. Neurosci. 1990;10:3421–3438. 11. Wu Y, Richard S, Parent A. The organization of the striatal output system: a single-cell juxtacellular labeling study in the rat. Neurosci Res. 2000;38(1):49–62. 12. Tepper JM, Bolam JP. Functional diversity and specificity of neostriatal interneurons. Curr Opin Neurobiol. 2004;14(6): 685–692. 13. Koos T, Tepper JM. Inhibitory control of neostriatal projection neurons by GABAergic interneurons. Nat Neurosci. 1999;2:467–472. 14. Prensa L, Parent A. The nigrostriatal pathway in the rat: a singleaxon study of the relationship between dorsal and ventral tier nigral neurons and the striosome/matrix striatal compartments. J Neurosci. 2001;21:7247–7260. 15. Cragg SJ, Baufreton J, Xue Y, Bolam JP, Bevan MD. Synaptic release of dopamine in the subthalamic nucleus. Eur J Neurosci. 2004;20(7):1788–1802. 16. Freund TF, Powell J, Smith AD. Tyrosine hydroxylase-immunoreactive boutons in synaptic contact with identified striatonigral neurons, with particular reference to dendritic spines. Neuroscience. 1984;13:1189–1215. 17. Pickel VM, Beckley SC, Joh TH, Reis DJ. Ultrastructural immunocytochemical localization of tyrosine hydroxylase in the neostriatum. Brain Res. 1981;225:373–385.

2.4: THE RELATIONSHIP BETWEEN DOPAMINERGIC AXONS AND GLUTAMATERGIC SYNAPSES 18. Groves PM, Linder JC, Young SJ. 5-Hydroxydopamine-labeled dopaminergic axons: three-dimensional reconstructions of axons, synapses and postsynaptic targets in rat neostriatum. Neuroscience. 1994;58:593–604. 19. Hanley JJ, Bolam JP. Synaptology of the nigrostriatal projection in relation to the compartmental organization of the neostriatum in the rat. Neuroscience. 1997;81(2):353–370. 20. Zahm DS. An electron microscopic morphometric comparison of tyrosine hydroxylase immunoreactive in the neostriatum and the nucleus accumbens core and shell. Brain Res. 1992;575:341–346. 21. Descarries L, Watkins KC, Garcia S, Bosler O, Doucet G. Dual character, asynaptic and synaptic, of the dopamine innervation in adult rat neostriatum: a quantitative autoradiographic and immunocytochemical analysis. J Comp Neurol. 1996;375(2): 167–186. 22. Smith Y, Bennett BD, Bolam JP, Parent A, Sadikot AF. Synaptic relationships between dopaminergic afferents and cortical or thalamic input in the sensorimotor territory of the striatum in monkey. J Comp Neurol. 1994;344:1–19. 23. Bouyer JJ, Park DH, Joh TH, Pickel VM. Chemical and structural analysis of the relation between cortical inputs and tyrosine hydroxylase–containing terminals in rat neostriatum. Brain Res. 1984;302:267–275. 24. Gerfen CR, Wilson CJ. The basal ganglia. In: Swanson LW, Bjo¨rklund A, Ho¨kfelt T, eds. Handbook of Chemical Neuroanatomy. Amsterdam: Elsevier; 1996:12:371–468. 25. Ragsdale CW, Graybiel AM. The fronto-striatal projection in the cat and monkey and its relationship to inhomogeneities established by acetyl-cholinesterase histochemistry. Brain Res. 1981;208:259–266. 26. Donoghue JP, Herkenham M. Neostriatal projections from individual cortical fields conform to histochemically distinct striatal compartments in the rat. Brain Res. 1986;397:397–403. 27. Malach R, Graybiel AM. Mosaic architecture of the somatic sensory-recipient sector of the cat’s striatum. J Neurosci. 1986;6:3436–3458. 28. Gerfen CR. The neostriatal mosaic: compartmentalization of corticostriatal input and striatonigral output systems. Nature. 1984;311:461–464. 29. Gerfen CR. The neostriatal mosaic: striatal patch-matrix organization is related to cortical lamination. Science. 1989;246:385–388. 30. Berendse HW, Galis-de-Graaf Y, Groenewegen HJ. Topographical organization and relationship with ventral striatal compartments of prefronto corticostriatal projections in the rat. J Comp Neurol. 1992;316:314–347. 31. Kincaid AE, Wilson CJ. Corticostriatal innervation of the patch and matrix in the rat neostriatum. J Comp Neurol. 1996; 374(4):578–592. 32. Graybiel AM. Network-level neuroplasticity in cortico-basal ganglia pathways. Parkinsonism Relat Disord. 2004; 10(5):293–296. 33. Flaherty AW, Graybiel AM. Input-output organization of the sensorimotor striatum in the squirrel monkey. J Neurosci. 1994;14:599–610. 34. Parthasarathy HB, Graybiel AM. Cortically driven immediateearly gene expression reflects modular influence of sensorimotor cortex on identified striatal neurons in the squirrel monkey. J. Neurosci. 1997;17(7):2477–2491. 35. Parthasarathy HB, Schall JD, Graybiel AM. Distributed but convergent ordering of corticostriatal projections: analysis of

36.

37.

38.

39.

40.

41.

42.

43.

44.

45.

46.

47.

48.

49.

50. 51.

52.

53.

57

the frontal eye field and the supplementary eye field in the macaque monkey. J. Neurosci. 1992;12:4468–4488. Smith Y, Bevan MD, Shink E, Bolam JP. Microcircuitry of the direct and indirect pathways of the basal ganglia. Neuroscience. 1998;86(2):353–387. Lacey CJ, Boyes J, Gerlach O, Chen L, Magill PJ, Bolam JP. GABA(B) receptors at glutamatergic synapses in the rat striatum. Neuroscience. 2005;136(4):1083–1095. Raju DV, Shah DJ, Wright TM, Hall RA, Smith Y. Differential synaptology of vGluT2-containing thalamostriatal afferents between the patch and matrix compartments in rats. J Comp Neurol. 2006;499(2):231–243. Fujiyama F, Unzai T, Nakamura K, Nomura S, Kaneko T. Difference in organization of corticostriatal and thalamostriatal synapses between patch and matrix compartments of rat neostriatum. Eur J Neurosci. 2006;24(10):2813–2824. Frotscher M, Rinne U, Hassler R, Wagner A. Termination of cortical afferents on identified neurons in the caudate nucleus of the cat. A combined Golgi-EM degeneration study. Exp Brain Res. 1981;41:329–337. Somogyi P, Bolam JP, Smith AD. Monosynaptic cortical input and local axon collaterals of identified striatonigral neurons. A light and electron microscopic study using the Golgi-peroxidase transport-degeneration procedure. J Comp Neurol. 1981;195:567–584. Kemp JM, Powell TPS. The structure of the caudate nucleus of the cat: light and electron microscopy. Philos Trans R Soc Lond. 1971;B 262:383–401. Hersch SM, Ciliax BJ, Gutekunst CA, et al. Electron microscopic analysis of D1 and D2 dopamine receptor proteins in the dorsal striatum and their synaptic relationships with motor corticostriatal afferents. J Neurosci. 1995;15:5222–5237. Lei W, Jiao Y, Del Mar N, Reiner A. Evidence for differential cortical input to direct pathway versus indirect pathway striatal projection neurons in rats. J Neurosci. 2004;24(38):8289–8299. Kincaid AE, Zheng T, Wilson CJ. Connectivity and convergence of single corticostriatal axons. J Neurosci. 1998;18(12):4722– 4731. Zheng T, Wilson CJ. Corticostriatal combinatorics: the implications of corticostriatal axonl arborizations. J Neurophysiol. 2002;87:1007–1017. Cowan RL, Wilson CJ, Emson PC, Heizmann CW. Parvalbumin-containing GABAergic interneurons in the rat neostriatum. J Comp Neurol. 1990;302:197–205. Kita H, Kosaka T, Heizmann CW. Parvalbumin-immunoreactive neurons in the rat neostriatum: a light and electron microscopic study. Brain Res. 1990;536:1–15. Kawaguchi Y. Physiological, morphological, and histochemical characterization of three classes of interneurons in rat neostriatum. J Neurosci. 1993;13:4908–4923. Tepper JM, Koos T, Wilson CJ. GABAergic microcircuits in the neostriatum. Trends Neurosci. 2004;27(11):662–669. Bennett BD, Bolam JP. Synaptic input and output of parvalbumin-immunoreactive neurones in the neostriatum of the rat. Neuroscience. 1994; 62:707–719. Lapper SR, Smith Y, Sadikot AF, Parent A, Bolam JP. Cortical input to parvalbumin-immunoreactive neurones in the putamen of the squirrel monkey. Brain Res. 1992;580: 215–224. Ramanathan S, Hanley JJ, Deniau J-M, Bolam JP. Synaptic convergence of motor and somatosensory cortical afferents onto GABAergic interneurons in the rat striatum. J Neurosci. 2002; 22:8158–8169.

58

NEUROANATOMY

54. Vuillet J, Kerkerian L, Kachidian P, Bosler O, Nioeullon A. Ultrastructural correlates of functional relationships between nigral dopaminergic or cortical afferent fibres and neuropeptide Y–containing neurons in the rat striatum. Neurosci Lett. 1989;100:99–104. 55. Wilson CJ, Chang HT, Kitai ST. Firing patterns and synaptic potentials of identified giant aspiny interneurons in the rat neostriatum. J Neurosci. 1990;10(2):508–519. 56. Reynolds JN, Wickens JR. The corticostriatal input to giant aspiny interneurons in the rat: a candidate pathway for synchronising the response to reward-related cues. Brain Res. 2004;1011(1):115–128. 57. Meredith GE, Wouterlood FG. Hippocampal and midline thalamic fibres and terminals in relation to the choline acetyltransferase–immunoreactive neurons in nucleus accumbens of the rat: a light and electron microscopic study. J Comp Neurol. 1990;296:204–221. 57a. Thomas TM, Smith Y, Levey AI, Hersch SM. Cortical inputs to m2-immunoreactive striatal interneurons in rat and monkey. Synapse. 2000; 37(4):252–261. 58. Smith Y, Raju DV, Pare JF, Sidibe´ M. The thalamostriatal system: a highly specific network of the basal ganglia circuitry. Trends Neurosci. 2004;27(9):520–527. 59. Groenewegen HJ, Berendse HW. The specificity of the ’nonspecific’ midline and intralaminar thalamic nuclei. Trends Neurosci. 1994;17(2):52–57. 60. Herkenham M, Pert CB. Mosaic distribution of opiate receptors, parafascicular projections and acetylcholinesterase in rat striatum. Nature. 1981;291:415–418. 61. Descheˆnes M, Bourassa J, Doan VD, Parent A. A single-cell study of the axonal projections arising from the posterior intralaminar thalamic nuclei in the rat. Eur J Neurosci. 1996;8(2):329–343. 62. Sadikot AF, Parent A, Smith Y, Bolam JP. Efferent connections of the centromedian and parafascicular nuclei in the squirrel monkey. A light and electron microscopic study of the thalamostriatal projection in relation to striatal heterogeneity. J Comp Neurol. 1992;320:228–242. 63. Ragsdale CW, Graybiel AM. Compartmental organization of the thalamostriatal connection in the cat. J Comp Neurol. 1991;311:134–167. 64. Descheˆnes M, Bourassa J, Parent A. Two different types of thalamic fibers innervate the rat striatum. Brain Res. 1995;701(1–2):288–292. 65. Lacey CJ, Bolam JP, Magill PJ. Novel and distinct operational principles of intralaminar thalamic neurons and their striatal projections. J Neurosci. 2007;27(16):4374–4384. 66. Xu ZC, Wilson CJ, Emson PC. Restoration of thalamostriatal projections in rat neostriatal grafts: An electron microscopic analysis. J Comp Neurol. 1991;303:22–34. 67. Lapper SR, Bolam JP. Input from the frontal cortex and the parafascicular nucleus to cholinergic interneurones in the dorsal striatum of the rat. Neuroscience. 1992;51:533–545. 68. Kemp JM, Powell TPS. The site of termination of afferent fibres in the caudate nucleus. Philos Trans R Soc Lond. 1971;B 262:413–427. 69. Sidibe´ M, Smith Y. Differential synaptic innervation of striatofugal neurones projecting to the internal or external segments of the globus pallidus by thalamic afferents in the squirrel monkey. J Comp Neurol. 1996;365(3):445–465. 70. Chung JW, Hassler R, Wagner A. Degeneration of two of nine types of synapses in the putamen after center median coagulation in the cat. Exp Brain Res. 1977;28:345–361.

71. Dube´ L, Smith AD, Bolam JP. Identification of synaptic terminals of thalamic or cortical origin in contact with distinct medium size spiny neurons in the rat neostriatum. J Comp Neurol. 1988; 267:455–471. 72. Raju DV, Ahern TH, Shah DJ, et al. Differential synaptic plasticity of the corticostriatal and thalamostriatal systems in an MPTP-treated monkey model of parkinsonism. Eur J Neurosci. 2008;27(7):1647–1658. 73. Sidibe´ M, Smith Y. Differential synaptic innervation of striatofugal neurones projecting to the internal or external segments of the globus pallidus by thalamic afferents in the squirrel monkey. J Comp Neurol. 1996;365(3):445–465. 74. Sidibe´ M, Smith Y. Thalamic inputs to striatal interneurons in monkeys: synaptic organization and co-localization of calcium binding proteins. Neuroscience. 1999;89:1189–1208. 75. Rudkin TM, Sadikot AF. Thalamic input to parvalbumin-immunoreactive GABAergic interneurons: organization in normal striatum and effect of neonatal decortication. Neuroscience. 1999;88(4):1165–1175. 76. Kachidian P, Vuillet J, Nieoullon A, Lafaille G, Goff LK-L. Striatal neuropeptide Y neurones are not a target for thalamic afferent fibres. Neuroreport. 1996;7(10):1665–1669. 77. Bolam JP, Bergman H, Graybiel A, et al. Molecules, microcircuits and motivated behaviour: microcircuits in the striatum. In: Grillner S, Graybiel A, eds. Microcircuits: The Interface between Neurons and Global Brain Function. Dahlem Workshop Report 93. Cambridge, MA: MIT Press; 2006:165–190. 78. Surmeier DJ, Ding J, Day M, Wang Z, Shen W. D1 and D2 dopamine-receptor modulation of striatal glutamatergic signaling in striatal medium spiny neurons. Trends Neurosci. 2007;30(5):228–235. 79. Calabresi P, Picconi B, Tozzi A, Di Filippo M. Dopaminemediated regulation of corticostriatal synaptic plasticity. Trends Neurosci. 2007;30(5):211–219. 80. Shen W, Flajolet M, Greengard P, Surmeier DJ. Dichotomous dopaminergic control of striatal synaptic plasticity. Science. 2008;321(5890):848–851. 81. Reynolds JN, Hyland BI, Wickens JR. A cellular mechanism of reward-related learning. Nature. 2001;413(6851):67–70. 82. Reynolds JN, Wickens JR. Substantia nigra dopamine regulates synaptic plasticity and membrane potential fluctuations in the rat neostriatum in vivo. Neuroscience. 2000;99(2):199–203. 83. Reynolds JN, Wickens JR. Dopamine-dependent plasticity of corticostriatal synapses. Neural Netw. 2002;15(4–6):507–521. 84. Wickens JR, Budd CS, Hyland BI, Arbuthnott GW. Striatal contributions to reward and decision making: making sense of regional variations in a reiterated processing matrix. Ann NY Acad Sci. 2007;1104:192–212. 85. Wickens JR, Reynolds JN, Hyland BI. Neural mechanisms of reward-related motor learning. Curr Opin Neurobiol. 2003;13(6):685–690. 86. Smeal RM, Gaspar RC, Keefe KA, Wilcox KS. A rat brain slice preparation for characterizing both thalamostriatal and corticostriatal afferents. J Neurosci Methods. 2007;159(2): 224–235. 87. Ding J, Peterson JD, Surmeier DJ. Corticostriatal and thalamostriatal synapses have distinctive properties. J Neurosci. 2008; 28(25):6483–6492. 88. Ingham CA, Hood SH, Taggart P, Arbuthnott GW. Plasticity of synapses in the rat neostriatum after unilateral lesion of the nigrostriatal dopaminergic pathway. J Neurosci. 1998; 18:4732–4743.

2.4: THE RELATIONSHIP BETWEEN DOPAMINERGIC AXONS AND GLUTAMATERGIC SYNAPSES 89. Day M, Wang Z, Ding J, et al. Selective elimination of glutamatergic synapses on striatopallidal neurons in Parkinson disease models. Nat Neurosci. 2006;9(2):251–259. 90. Ingham CA, Hood SH, Arbuthnott GW. Spine density on neostriatal neurones changes with 6-hydroxydopamine lesions and with age. Brain Res. 1989;503:334–338. 91. Smith AD, Bolam JP. The neural network of the basal ganglia as revealed by the study of synaptic connections of identified neurones. Trends Neurosci. 1990;13:259–265. 92. Pinto A, Jankowski M, Sesack SR. Projections from the paraventricular nucleus of the thalamus to the rat prefrontal cortex and nucleus accumbens shell: ultrastructural characteristics and spatial relationships with dopamine afferents. J Comp Neurol. 2003;459(2):142–155. 93. Goldman-Rakic PS, Leranth C, Williams SM, Mons N, Geffard M. Dopamine synaptic complex with pyramidal neurons in primate cerebral cortex. Proc Natl Acad Sci USA. 1989;86:9015–9019. 94. Pinto A, Sesack SR. Ultrastructural analysis of prefrontal cortical inputs to the rat amygdala: spatial relationships to presumed dopamine axons and D1 and D2 receptors. Brain Struct Funct. 2008;213(1–2):159–175. 95. Ni B, Wu X, Yan GM, Wang J, Paul SM. Regional expression and cellular localization of the Na(+)-dependent inorganic phosphate cotransporter of rat brain. J Neurosci. 1995;15(8): 5789–5799. 96. Otis TS, Kavanaugh MP. Isolation of current components and partial reaction cycles in the glial glutamate transporter EAAT2. J Neurosci. 2000;20(8):2749–2757. 97. Bellocchio EE, Reimer RJ, Fremeau RT Jr, Edwards RH. Uptake of glutamate into synaptic vesicles by an inorganic phosphate transporter. Science. 2000;289(5481):957–960. 98. Fremeau RT Jr, Troyer MD, Pahner I, et al. The expression of vesicular glutamate transporters defines two classes of excitatory synapse. Neuron. 2001;31(2):247–260. 99. Fujiyama F, Kuramoto E, Okamoto K, et al. Presynaptic localization of an AMPA-type glutamate receptor in corticostriatal and

59

thalamostriatal axon terminals. Eur J Neurosci. 2004; 20(12):3322–3330. 100. Barroso-Chinea P, Castle M, Aymerich MS, Lanciego JL. Expression of vesicular glutamate transporters 1 and 2 in the cells of origin of the rat thalamostriatal pathway. J Chem Neuroanat. 2008;35(1):101–107. 101. Barroso-Chinea P, Castle M, Aymerich MS, et al. Expression of the mRNAs encoding for the vesicular glutamate transporters 1 and 2 in the rat thalamus. J Comp Neurol. 2007;501(5):703–715. 102. Agnati LF, Zoli M, Stromberg I, Fuxe K. Intercellular communication in the brain: wiring versus volume transmission. Neuroscience. 1995;69(3):711–726. 103. Rice ME, Cragg SJ. Dopamine spillover after quantal release: rethinking dopamine transmission in the nigrostriatal pathway. Brain Res Rev. 2008;58(2):303–313. 104. Cragg SJ, Rice ME. DAncing past the DAT at a DA synapse. Trends Neurosci. 2004;27(5):270–277. 105. Arbuthnott GW, Wickens J. Space, time and dopamine. Trends Neurosci. 2007;30(2):62–69. 106. Gonon F. Prolonged and extrasynaptic excitatory action of dopamine mediated by D1 receptors in the rat striatum in vivo. J Neurosci. 1997;17(15):5972–5978. 107. Delle Donne KT, Sesack SR, Pickel VM. Ultrastructural immunocytochemical localization of the dopamine D2 receptor within GABAergic neurons of the rat striatum. Brain Res. 1997;746(1–2):239–255. 108. Sesack SR, Aoki C, Pickel VM. Ultrastructural localization of D2 receptor-like immunoreactivity in midbrain dopamine neurons and their striatal targets. J Neurosci. 1994; 14(1):88–106. 109. Yung KKL, Bolam JP, Smith AD, Hersch SM, Ciliax BJ, Levey AI. Immunocytochemical localization of D1 and D2 dopamine receptors in the basal ganglia of the rat: light and electron microscopy. Neuroscience. 1995;65:709–730.

This page intentionally left blank

3 Molecular pharmacology

This page intentionally left blank

3.1

Molecular Pharmacology of the Dopamine Receptors MI CH ELE L . R ANK IN , LIS A A . HA ZELWO OD, R . BEN JA MI N F REE , YO ON NA MK UNG, ELIZ AB E TH B. RE X , REB ECC A A . RO OF, AND DAVI D R. S I B LE Y

INTRODUCTION Dopamine receptors are rhodopsin-like seven-transmembrane receptors (also called G protein–coupled receptors) that mediate the central and peripheral actions of dopamine. Dopamine receptors are most abundant in pituitary and brain, particularly in the basal forebrain, but are also found in the retina and in peripheral organs such as the kidney. Stimulation of dopamine receptors modulates natriuresis in the kidney, as well as cell division and hormone synthesis and secretion in the pituitary. Brain dopamine receptors regulate movement and locomotion, motivation, and working memory. Five subtypes of mammalian dopamine receptors have been identified that are divided into D1-like (D1, D5) or D2-like (D2, D3, D4) subgroups. The D1-like receptors couple primarily to the Gs family of G proteins (Gs and Golf), whereas the D2-like receptors couple primarily to the Gi/o family. This chapter covers the molecular pharmacology of the five dopamine receptor subtypes.

THE D1DOPAMINE RECEPTOR SUBTYPE The D1 dopamine receptor (D1DAR; Fig. 3.1.1) subtype (also called the D1A subtype in some literature) belongs to the D1-like family of dopamine receptors, and it is the discovery and characterization of this receptor that gives this family its name.1–3 The D1DAR (GenBank Accession NP_000785; located on chromosome 5q35.1) exhibits the most conserved sequence of all the DARs, featuring an extended C-terminus and a shortened third intracellular loop (ICL3) when compared to the D2like receptor family. Of all the individual DAR knockout mice, the D1DAR knockout exhibits the most severe phenotypes, including spatial learning deficits,4 hyperactivity,5 and abnormal memory retention,6 emphasizing the functional importance of this receptor subtype. The

D5DAR (also called D1B) is the remaining receptor subtype that comprises the D1-like family and will be discussed in a later section of this chapter. D1DAR Structure The D1DAR belongs to the class A or rhodopsin family of G protein–coupled receptors (GPCRs).7 The past couple of years have yielded some exciting advances in the elucidation of the structural topography of GPCRs. Although GPCRs are notoriously difficult to crystallize, there are now three crystal structures available for this class of receptors: rhodopsin,8 the b1-adrenergic receptor,9 and the b2-adrenergic receptor.10,11 The majority of the data gleaned from these structural studies reveals the details of the intramembrane helical interfaces, whereas the more mobile intracellular loops and carboxyl tail region are too disordered to discern. These data provide insight into how ligand binding occurs and alters the protein structure to transduce the activated receptor conformation to activation of G protein and downstream signaling events.12,13 Given that all clinically used antipsychotic drugs bind to and antagonize dopamine receptors,14 structural studies will continue to play a major role in research and development for improved therapies that involve the dopaminergic signaling system. Posttranslational modifications Many GPCRs have been shown to be palmitoylated at cysteine residues in the carboxyl tail region. This covalent modification results in the formation of a fourth intracellular loop that is folded into an amphiphilic a-helix that lies parallel to the intracellular surface of the plasma membrane.15,16 Because palmitoylation is a reversible modification and can provide a mechanism for membrane localization, it has been proposed to participate in membrane association and/or trafficking.16 In addition, palmitoylation 63

64

MOLECULAR PHARMACOLOGY

FIGURE 3.1.1. Structure of the rat D1DAR. The figure shows the locations of amino acid residues associated with several functional features. Glycosylation of Asn residues is indicated by ‘‘–CHO’’.20 Ser256, Ser258, and Ser259 are essential for efficient arrestin association.40 Thr268 is associated with the rate of desensitization and receptor trafficking.111,269 Phe313 and Trp318, located in TMD7, are essential for caveolaemediated endocytosis.35 Cys347 and Cys351 are palmitoylated.19 The region within the carboxyl tail represented by gray residues is essential for efficient endocytic recycling.270 Thr428 and Ser431 are constitutively phosphorylated by GRK4a.72 See text for details.

of the b2-adrenergic receptor has been shown to govern the accessibility of a protein kinase A (PKA) phosphorylation site in the carboxyl tail that affects desensitization of this receptor.17,18 The D1DAR is palmitoylated on Cys347 and Cys351 in the proximal portion of the carboxyl tail.19 The functional significance of this modification remains unknown. Efficient plasma membrane localization of several GPCRs has been shown to be dependent on N-linked glycosylation. Indeed, the D1DAR is glycosylated at Asn5 and Asn175; however, chemical or mutational inhibition of glycosylation of the D1DAR has no effect on proper targeting of the receptor to the cell surface. In contrast, glycosylation of the D5DAR is essential to proper cell surface trafficking and concomitant ligand binding.20 Functional domains of D1DARs Numerous studies have been performed on the D1-like and D2-like receptors using a combination of mutational

strategies and pharmacological approaches to correlate structural domains within receptor proteins to function and subtype specificity. Studies focused on D1-like receptors utilized D1/D5 chimeras to show that the C-terminus of the D1DAR imparts both lower dopamine affinity and reduced constitutive activity to this receptor subtype when compared to that of the D5DAR, while the third extracellular loop (ECL3) mediates reduced D1DAR dopamine potency compared to the D5DAR.21,22 Furthermore, Sugamori et al.23 have shown that the C-terminus of the D1DAR is responsible for the actions of benzazepines—antagonists for D1 and partial agonists for D5. In a separate set of experiments, chimeras were designed to study ligand binding and adenylyl cyclase (AC) activation properties associated with D1-like and D2-like receptors. To accomplish this, a chimera composed of the entire proximal portion of the D1DAR extending from the N-terminus to ICL3 was fused to the distal portion of the D2DAR extending from transmembrane domain 6(TMD6) through the C-terminus. Stimulation of cells expressing this chimera with

3.1: MOLECULAR PHARMACOLOGY OF THE DOPAMINE RECEPTORS

a D2DAR-selective ligand resulted in activation of AC (like D1DARs), indicating that the portion of the receptor imparting D2-selective ligand specificity resides within TMD6 and TMD7.24 Early work on the b-adrenergic receptor revealed that the ligand-binding pocket in catecholamine receptors is comprised of TMD3, TMD5, and TMD6, and that Asp102 and a cluster of Ser residues are particularly important.25 A naturally occurring polymorphism at S199A in TMD5 results in a decreased affinity for the antagonist SCH-23390.26 More recent site-directed mutagenesis experiments illustrate that Trp99 and Ala195 in D1DAR are important for the interaction with D1-selective, rather then D3-selective, ligands.27 Higher order structures D1DAR interacts with numerous other proteins to form higher order structures including other GPCRs to form both homo-oligomers28 and hetero-oligomers, such as those observed with D2DAR29 and A1 adenosine receptors30. Many GPCRs also form complexes with ion channels and transporters including the sodium, potassium adenosine triphosphatase (Na+, K+-ATPase)31 and N-methyl-d-aspartate (NMDA) receptors32; adaptors/ trafficking proteins such as DRiP78,33 calnexin,34 caveolin,35 N-ethylmaleimide-sensitive factor,36 and sorting nexin-136; regulatory proteins such as G proteins and arrestins37; and kinases.38 In general, ICL2, ICL3, and the C-terminus of GPCRs are responsible for G protein coupling,39 while b-arrestin binding to D1DAR is mediated via ICL3 interactions.40

65

treat PD clinically are as effective as levodopa, the agent that comes closest is apomorphine—an agonist that exhibits D1 selectivity.45 It is generally agreed that D2 antagonists are effective antipsychotics; however, D1 agonists may also be useful for treatment of the cognitive impairment associated with schizophrenia and may even be neuroprotective.45 MPTP (1-methy-4-phenyl-1,2,3,6-tetrahydropyridine) is a neurotoxin that causes permanent symptoms of PD by killing certain neurons in the substantia nigra of the brain. One study has shown that the full D1DAR agonist dihydrexidine improved cognitive function in MPTP-treated monkeys,46 while partial D1 agonists and D2 agonists demonstrated little effect.47 Ligand structures D1DAR ligands can be divided into separate structural classes (Tables 3.1.1 and 3.1.2). The first D1DAR ligands developed were phenyltetrahydrobenzazepines (such as SCH-23390 and SKF-83959), followed by the development of rigid analogs of b-phenyldopamines (such as dihydrexidine). More recently developed D1selective ligands include constrained phenylbenzazepines and polycyclic analogues of dihydrexidine.48 Although few have made it to the market yet, D1DAR ligands have been developed and used in clinical trials for the treatment of drug abuse, sleep disorders, obesity, PD, and schizophrenia.48 Some of the most common agonists and antagonists for the D1DAR are listed in Tables 3.1.1 and 3.1.2, respectively.

D1DAR Signaling Mechanisms D1DAR Pharmacology and Localization in the Brain Therapeutic potential The D1DAR is the most highly expressed DAR and is localized within the forebrain in areas such as the caudate putamen, substantia nigra, nucleus accumbens, hypothalamus, frontal cortex, and olfactory bulb. As such, it is implicated in cognitive and motor functions, as well as substance abuse, Parkinson’s disease (PD), and schizophrenia. Although the D2DAR has been the most frequently targeted DAR for the treatment of PD, there are data to suggest that D1DAR agonists may be beneficial. For example, dihydrexidine41 and other full agonists42,43 have demonstrated antiparkinson actions in animals, although it should be noted that dihydrexidine is only 10-fold more selective for D1 than for D2; thus, D2 activity could account for some of the antiparkinson activity.44 While no DAR agonists currently used to

G protein coupling D1DAR couples to heterotrimeric (a, b, g) GTP-binding proteins (G proteins) that activate AC, resulting in an accumulation of the second messenger cyclic adenosine 30 ,50 -cyclic monophosphate (cAMP) and a concomitant activation of cAMP-dependent PKA. The domains of the D1DAR involved in G protein interaction have been mapped to ECL2 and ECL3, as well as to the proximal portion of the carboxyl tail.49 The specific G proteins that couple to the D1DAR are determined by the tissue where the receptor is expressed. The prototypical Ga protein associated with AC activation is Gas. This G protein subunit is ubiquitously expressed in a variety of tissues and cell culture lines.50 D1DAR signaling has been studied in a multitude of cell types and displays robust coupling to Gas upon agonist activation, indicating that D1DAR signaling is effectively mediated by this G protein50; however, the

66

MOLECULAR PHARMACOLOGY

TABLE 3.1.1.

D1DAR Agonists

D1 Agonist

Structural Class

Ki (nM) 2340

Dopamine

Catecholamine

(þ)SKF-82526 Fendoldopam

Benzazepine

17

NPA

Nonergoline

SKF-38393 Dihydrexidine

Other Targets

Therapeutic/Experimental Use

All DAR, D5 > D1 also aAR

Hemodynamic imbalances

D1- and D2-like

Hypertension

1816

D2-like > D1-like

Experimental tool

Benzazepine

150

D1-like > D2-like

Experimental tool

Benzazepine

33

D1-like  D2-like; also aAR

Clinical trial for cocaine disorders

Note: Some common D1 agonists are listed with Ki values and relative affinities for non-D1 targets from the NIMH Psychoactive Drug Screening Program (PDSP) database.268 Abreviations: AR, adrenergic receptors; NPA, N-propylnorapomorphine.

TABLE 3.1.2.

D1DAR Antagonists

D1 Antagonist

Structural Class

Ki (nM)

SCH-23390

Benzazepine

0.35

D1-like > 5-HT > D2-like

Experimental tool

SCH-39166 Ecopipam

Benzazepine

3.6

D1-like > D2-like

Experimental tool

SKF-83566

Benzazepine

0.3

D1-like > 5-HT

Experimental tool

Thioridazine

Phenothiazine

All DARs, 5-HT, aAR, H1

Antidepressant, antianxiety, antipsychotic

100

Other Targets

Therapeutic/Experimental Use

Chlorpromazine

phenothiazine

73

All DARs, 5-HT, MR, AR

Antipsychotic, tranquilizer, antiemetic

Fluphenazine

Phenothiazine

21

All DARs, 5-HT, MR, AR, HT

Antipsychotic

Note: Some common D1 antagonists are listed with Ki values and relative affinities for non-D1 targets from the NIMH Psychoactive Drug Screening Program (PDSP) database. Abbreviations: AR, adrenergic receptors; 5-HT, serotonin receptors; MR, muscarinic receptors; HT, histamine receptors.268

neostriatum of the brain is the region where the D1DAR is most abundantly expressed, yet this region lacks robust Gas expression. Gaolf, on the other hand, is abundantly expressed in the neostriatum and couples positively to activation of AC.51 Support for the theory that D1DAR can effectively couple to Gaolf was demonstrated by Gaolf knockout mice that displayed deficient D1DAR-mediated behaviors.51 There is little data identifying the specific b and g G protein subunits that participate in D1DAR signaling; however, in HEK293 tissue culture cells, depletion of g7 reduces D1DAR-mediated AC stimulation and decreases the abundance of the b1 subunit.52 These data suggest that in HEK293 cells the heterotrimeric G proteins that mediate D1DAR signaling are Gasb1g7. D1DAR coupling to Gaq remains a controversial topic due to the inability to reconcile contradictory observations reported by several independent research groups. Gaq-mediated signal transduction generally involves activation of phospholipase C (PLC) that cleaves phosphatidylinositol-4,5-bisphosphate (PIP2), producing the second messenger diacylglycerol (DAG), that activates several forms of protein kinase C (PKC), and inositol-1,4,5-trisphosphate (IP3) that induces release of calcium from intracellular stores through

binding of IP3 receptors (IP3Rs) located on the endoplasmic reticulum. It has been proposed that there exists a new member of the D1-like family of dopamine receptors that couples to Gq signaling pathways but has yet to be identified. Support for this theory stems from the observations that in D1DAR null mice, cAMP signaling and behavior are greatly diminished with the use of D1DAR agonists, while Gq signaling pathways remain intact, as does coimmunoprecipitation of [3H]SCH23390 binding sites with Gaq protein.53,54 Others disagree with the notion that there exists an unidentified member of the D1-like DARs and propose that Gq signaling represents an alternative signaling pathway for D1DAR.55 Sahu et al.56 have looked at IP3 accumulation and DAG production in D5DAR knockout mice and show that accumulation of these second messengers is severely impaired in several brain tissues, diminishing the role of the D1DAR receptor in this pathway. Recent evidence also suggests that Gq coupling may be accomplished by hetero-oligomerization of D1DARs with D2DARs, generating a multi-receptor complex that exhibits altered G protein coupling and requires activation of each receptor in the complex to effect Gq activation.57 Oligomerization of different receptor proteins within and across G

3.1: MOLECULAR PHARMACOLOGY OF THE DOPAMINE RECEPTORS

protein–coupled receptor (GPCR) classes to form signaling units with unique pharmacology is an exciting topic in GPCR research; however, it is beyond the scope of this discussion. D1DAR signaling through cAMP and PKA Stimulation of D1DAR coupled to Gas/olf pathways results in the activation of AC, production of cAMP, and concomitant activation of PKA. This sequence of events represents the most widely studied signaling paradigm for the D1DAR subtype. The activation of PKA through this signaling paradigm modulates a plethora of downstream targets that contribute to the overall D1DAR response. One of the most widely studied substrates for activated PKA mediated by dopamine is DARPP-32 (dopamine- and cAMP-regulated phosphoprotein, 32 kDa).58 DARPP-32 is enriched in the neostriatum and, once activated, amplifies D1DAR signaling by both preventing inactivation of PKA and preventing dephosphorylation, and hence inhibition, of many of the downstream targets of PKA. DARPP-32 is only mentioned briefly here and will be discussed in more detail in later chapters of this volume. D1DAR-PKA regulation of ion channels Here, D1DAR regulation of ion channels will only be discussed briefly. For a more in-depth review of DAR regulation of ion channels, the reader is directed to an excellent review by Neve et al.38 D1DAR regulation of ion channels is achieved by activation of PKA that results in the direct phosphorylation of ion channel subunits and other cellular effectors to alter channel activity. DARPP-32 phosphorylation resulting from D1DAR activation inhibits protein phosphatase 1 (PP1), that in turn enhances phosphorylation of the sodium channel at Ser573 to decrease sodium channel activity.59,60 D1DAR activation of PKA also decreases inwardly rectifying potassium channels and N- and P/Q-type calcium channels. Conversely, PKA activation increases L-type calcium channels and NMDA and a-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) currents, as well as modulating GABA currents (reviewed in Neve et al.38). D1DAR-PKA activation of CREB The signal transduction events that have been discussed so far represent instantaneous signaling events that occur upon initial D1DAR activation. Sustained receptor occupancy that exists during addictive

67

behaviors or therapeutic treatment induces long-term changes in signaling pathways mediated by altered gene expression that result in modifications of neural networks that are translated into behavioral responses. D1DAR activation of PKA results in phosphorylation of the cAMP response element binding protein (CREB) by PKA on Ser133 of CREB. Phosphorylated CREB dimerizes and acts as a transcription factor by binding to cAMP response element (CRE) DNA sequences located in the upstream promoter regions of a variety of genes, including the immediate early gene family of transcription factors Fos and Jun. Modulation of DAR activity leads to the expression of several neurotransmitter genes such as the those of the neuropeptides enkephalin and neurotensin (discussed in Adams et al.14).

D1DAR Regulation Desensitization Upon agonist activation, GPCRs undergo desensitization, a process that results in a waning of receptor response under continued agonist stimulation. Desensitization involves phosphorylation of the receptor by G protein receptor kinases (GRKs) and/or second messenger kinases such as PKA or PKC. Phosphorylation of GPCRs is categorized as either homologous or heterologous, with most GPCRs undergoing both types of phosphorylation. Heterologous phosphorylation occurs when a GPCR becomes phosphorylated by kinases activated by a separate signaling pathway. Kinases mediating heterologous desensitization are second-messenger-activated kinases such as PKA and PKC. Homologous desensitization of GPCRs results in the phosphorylation of the receptor by kinases activated as a result of the signaling events initiated upon activation of that receptor. Homologous phosphorylation of GPCRs is primarily mediated by GRKs. A general model of desensitization involves arrestin binding to phosphorylated receptor and concomitant uncoupling of the receptor from G protein, abrogating second messenger production. The arrestin-bound receptor is then internalized via clathrin-coated pits. Once internalized, the receptors are either dephosphorylated by phosphatases within the endocytic vesicle and recycled to the plasma membrane for additional signaling (resensitization) or targeted to lysosomal vesicles for degradation (down regulation). Recent studies have shown that this model for GPCR desensitization contains many layers of complexity, with individual receptors displaying unique modes of

68

MOLECULAR PHARMACOLOGY

regulation, abolishing a universal theme for GPCR desensitization.61,62 To date, D1DAR desensitization has been studied using a variety of systems including tissue sections, primary cell cultures, cell cultures expressing endogenous receptor, and heterologous expression of receptors both stably and transiently in various cell culture lines.63,64 Since the studies were performed in a variety of cellular environments, not all of the data concerning D1DAR regulation are consistent; however, several lines of evidence exist that link D1DAR phosphorylation to desensitization. Gardner et al.65 showed that the potency for dopamine to induce D1DAR phosphorylation (EC50 ~200 nM) was identical to that required to initiate desensitization. Furthermore, D1DAR phosphorylation (t1/2